Conversion of Gaussian to 51 Units SI
Gaussian
Fundamental
Ratio·
quantity
SI Symbol
Units
Units
Symbol
Gaussian
10)/41( 4l1:xlO-'
Magnetization
4,M
Permeability of free space
-
Anisotropy Exchange Gyromagnetic ratio Gilbert parameter
K
,A •
-
•
-
Parameter
Defining
Units
Defining
Units
Ratio·
Magnetic flekl H strength Magnetic induction B Flux Energy product Demagnetizatjon N factor Volume
•
,.K
susexptibility Permeability Anisotropy field Stability factor
Wall energy density Characteristic length Wall mobility
m meter cm centimeter
M
Aim
J.lo",4IlX
10- 7
erg/em J K erg/em A (S·Oe)-1 Y
Wb/Am
10- 1 10-)
JIm)
Jim (s-A/ml-l 411'/10 J I
(ormula
formula
IH-dl.4IfNI/1O De
JH,-dl=NI
Aim
IOJ/4J11'
Wblm J Wb
10-· 10-' 10- 1
B_H+41tM
G
B=1JJ.H+M)
.-fB·dA BH/8lf.=-MHfl
M,
.~fB.dA
Jim' -
H. __ NM
crgfcm'
BHf1.=:>JloMHfl
-
H.=-NM
l-dM/dH
-
,.=dM/dH
-
4K
-
lJ.=dB/dH
-
411 X 10- 7
10'/411:
Jl-dB/dH -1+4n:X H, Hr -2K/M Q Q_H./4rtM P
•
lJ_4j.!AK
I
l_a/4nM2
o -Y~ P~ P-w· « K
,
G
,-
second
N number of turns
1/4/f
=JlJI+K)
-Po
Q=H.IM
Aim -
ergfcm2
a=4VA'K
J/m 2
10- 3
om
l=alp oM 2
m
10- 2
m 2 /s·A
4nxlO- l
De
-
cm/s·Oe
A Ampere J Joule
Hr =2KIJloM
~a.= 1~ • K 0, Oersted
G
Wb Weber
M,
1
Gauss Maxwell
• To obtain values in 51 units multiply value in Gaussian units by the respective ratio
The effect of Ul",U- 1 for various U's and for a = X,Y,Z. The bottom line applies for spin! only 1.
I,
1,
U e i91 %
1%
I,oos8-/~sin8
I~
ei(Jl y
/%cos8+I~sin8
I,
/~
ei91~
/% cos 8 -I, sin 8 1. c05(8/2)
1,0058+1% sin 8
I,
e-i(JI~S~
cos 8+1, sin 8 cos8-/% sin 8
1, 1, 000(812) + 1,(25, 5;0(812)) \ • • - 1.(25, ';0(p/2))
.
,
'.' Useful relations for unitary operators U. (U is a function of I, or S, or I and S)
Uj,(J,5lf,U,5W-' = Uf,U,5jU-'Uf,U,5)U-' Uexp[i/U,S)IU- 1 = exp(iU I(J, S}U-lj For spin!:
.
" ". 1 - '",
,;1.
ccMi·(I~B~,=·cos (8/2)
sin
(I~9)
12 - 1 "'-'I
=
2l~ sin
f
,-
-,
.
(8/2)
"a~:z:y:;\''''' -"
:~
,..,
1%111 = t/~ I'. : lL.DltliI.cyclic permutations "7 z r->',_ .. 1, 1% = -'2
,.
..
,
.
..
"
;,
...
:1 ',' -
,
.
.
, ., .
..
1
Springer Series in Solid-State Sciences Edited by Peler Fuldc
c. P. Slichter
Springer Series in Solid-State Sciences Editors: M. Cardona
P. Fuldc
K. von Klitzing
Managing Editor: H. K. V. Lotsch so
l\1ull~ Diffnl<1io.
orX·Ra)sll. CryMloIs
Vollllnt$ 1-49 a'C' listed at
70
By Shih·Lin Chan, 51 ",",non ScallerinC ill Cond~lOH
Ed;lon:
w. Eisc:mncngcr.l<. La6mann.
and S. 06Uinger
H.·J. Queisscr l~
71 lIilh Mag.etl~ Adds in Semirondudor I'hlsKs Edilor: G. Landwehr
52 S.....,..,unduel;"ily In Macnelk .nd hu.;" Mareri.ls EdilOr5: T. MalS\lb.,a ;",0.1 fl. KOlan;
72 One·l)imenslonal Condu(Iors
53 '1"'u_l}imcn,iunal SyaIN"" II.ICr(lllrUcW,cs. Ind Supc,lalli...,.
73 Qnantum Solid·Stale l'h,S''''
1:0.1;11'''': G. Bauer, F. Kuchar. and U. Ikinrich 54 Macne';" Excilalions Ind }ludu.tio", Editors: S. l.o\'~Y. U. Ball/Cl&ni, F. Ikm:a. .nd V. T"I""ui 55 1HllIeo
Ed,tor. G. Harlxke S8 The Renlrsion Melhod ..... Itt Applinlliu.s Edit",,: D. Pellifof and D. Weai •• 59 1),'II'I\,ind V _ .nd Ord~.;"1 Oft Solid Sprfa~n Edilors: A. Yoshimori and M, Tsukada 60 ~;~dtonic l'rMesses In SoUds By M. Ucta. H. KanUlki, K. Kobayasbi. Y. To)'o..wa. and E. Hanamo.a 61 l.ocali..lion, Inl~n.tlion, udTnnsporl I'henDlMna Edilon: 8. Kramer.G. lkrgmann, and Y. Uruynscncc:k 62 ll1rory .... 11~...y t"enrWo-.J .IId "a~~ Hun••l....... Edilon: T. Kasu)~ and T. Sa50 63
6J
6S
66
67
68 (R
t-:Ia1""
Properlies of pol)'.......-.nd R....lH Com,-..... Edilors: H. Kuzmany, M. M~hrin'. and S. Roeh S)'mmetnni. PIo) ..... GroupThcory .... ~ 10 Physical Problems By W.l.udYoiland C. Fallu Phonon<: TMory .nd £lo:perlmulS II Experimentsand Inlerp.etation of E~pt:rimemal Resuhs By I'. Uriiesch I'honon" Tho:<>ry .nd Experlmenls III Phenomena Relaled 10 Phonon, Lly I'. IIrliesch l'wu.Dil1len,iOl,al S)'Slen,s: I'h)"oin .nd N~.. D~"ir" Editors: G. Bauer. F. Kuehal. and H. H~inrich I'honon SnIt1erinlln CondeHed Mall~. V Ed'iors: A.C. Anderson andJ. P. Wolle Nonlinuril)' in eonok-" Malle. Ed,lors: A. R. Bishop, D. K. Campbell. p. Kuma•• and S.E. Trul1;nlC'
md of the boot
llamillonlaaS!o P1l_l>ial:nms "The EJ«lronir and Slatistical·Mechanical Theory ofsp-8ondcd M~I.kand Allo)' By J. 'h'""r ~'.-
lIyS. Kagosh;m •. H. NagaSllwa. andT. Sambollgi Edilors: S. V. Vonsovsky and M. I. Katsnebon 74 QUlntum Monic Carlu Mdhoch;n Equilibrium and Nonequilibrium Syslems Editor: M. Su~uki
7S
E1«1roni~ SIructU'~ ..d Oplinll I'ropt,Iiro; 01 Seraicondl>etors ByM.L.ColKnandJ R.Cbclit""u)·
16 EJectroaic Properlies oIC_jorpttil Poll'MIS Edilors: H. KlaZmlny. M. M~hrins. and S. R<Mh
Principles of Magnetic Resonance Third Enlarged and Updated Edition
11 FenniSoorfatt Etrec'b
Ed,lors: J. Kondoand A. Yosllimori 78 <; .....1' Theory ..... Ib Applkaliom iro P")!oio DyT. Inui. Y. Tana~. and Y. On<.>c:kn 19
F.I~m~nl"y £lo:riCaliool. in Quntum .1uilk Edilors: K. Ohl>
80
Monl~
~l
C"re·L~,d SpeWO!iC<JPl' In C"ndensed Editors: J. Kan"n'ori and A. KOlani
112
With 185 Figures
CarloSin,ull,i"" in Slalislical Ph'·"n An Inuoouelion By K. Binder and I). W. HeermmlO
In,.odu~,i"n
S)'~lems
"' Ph"l...,n,i"i"n SpcclroKOpy
By S. Hurner 8J Ph)~ lod 1'«hnolo&>'''''S"bmi
roce:ssa By E. Fd Ind G. Sa.uer~nn 87 II .... Macnelk ndd.inSemirondlKl'" P")sin II Edilor: G. I.. ndweh, 88 OrgankSupc,coI,du'ily Editors: J.G.lkdnonand K.A. Miil~. 91 tl«Ironoior ..... pema 01 ConjlOpiH poI)-.rterS III Qasic Models and ....pplnlions EditOD: H. Kuzmany. M. Mehrinl. and S. ROIh
Springer-Verlag Berlin Heidelberg New York London Paris Tokyo Hong Kong
Preface to the Third Edition
The first edition of Ihis book was wrilten in 1961 when I was Morris Loeb LeclUrer in Physics 31 Harvard. In the preface I wrote: "TIle problem faced by a beginner today is enonnous. If he attempts to read a current aniele. he oftcn finds that the first paragraph refers to an earlier paper on which the whole article is based. and with which the author naturally assumes familiarity. That reference in turn is based on another, so the hapless student finds himself in a seemingly endless retrcat. I have felt that graduate siudems or others beginning research in magnetic resonance needed a book which really went into the details of ca1culUlions, yet was aimed al the beginner rather than the expert." The original gool was to treat only those topics that are essential to an understanding of the literature. Thus the goal was to be selective rather than comprehensive. With the passage of time, important new concepts were becoming SO all-pervasive that I felt the need to add them. That led to the second edition, which Dr. Latsch, Physics Editor of Springer~Verlag,encouraged me to write and which helped launch the Springer Series in Solid-State Sciences. Now, ten years later, that book (and its 1980 revised printing) is no longer available. Meanwhile, workers in magnetic resonance have continued to develop startling new insights. There are new topics which are so important that they must be included in a book that is intended to be an intrOductory text. The original course was one semester - the book had more topics than I taught in that amount of time. The new edition clearly would need a full year. Throughout the history of this book, I have considered it as a textbook. Thus the aim was to explain in a rigorous but physical manner concepts which are essential for the student of magnetic resonance. Rather than giving an exhaustive treatment of anyone topic, my intention has been to help prepare a student to read the literature on that topic. The main additions include an enlargement and modest rewrite of the topic of double resonance, explanations of one- and two-dimensional Fourier tmnsfonn methods, of coherence transfer, of multiple quantum coherence, and of important topics related to dipolar coupling that underlie the method of spin-flip line narrowing. In the earlier editions, the chapter on the density matrix focused almost exclusively on its use in analyzing relaxation processes. The new edition explains its use in analyzing the effect of rf fields.
v
The first edition of this book came oot shortly after the publication of Anatole Abragam's classic-to-be Principles of Nuclear Magnetic Resonance. There were not many other books on the market Today there are many books on magnetic resonance, a large fraction written by some of the most illustrious founders of the field and practitioners of the art. The styles of these books are as varied as the scientific styles of their authors. As I have explored these marvellous books, I have been struck by the opportunity that exists today to get the flavor of the scientific thought of these great scientists. When I wrote the first edition of this book at Harvard, I had no duties other than preparing and giving the lectures. The third edition has been written on evenings and weekends while I tried to carry on my nonnal activities. It would not have been possible without the expert, dedicated. and always cheerful support of my secretary, Ann Wells. Her association with the book has been long, since she helped type the original edition. I am also grateful for the help of Jamie Froman in the final stages. Urbana, Illinois June 1989
Charles P. SUchter
Contents
1.
Elements of Resonance .........................•.......... 1.1 Introduction ...................................•..•. 1.2 Simple Resonance Theory ........................• _••. 1.3 Absorption of Energy and Spin-Lanice Relaxation . Theory . Motion of Isolated Spins-Classical Treattnent . Quantum Mechanical Description of Spin in a Static Field Equations of Motion of the Expectation Value . Effect of Alternating Magnetic Fields . Exponential Operators . Quantum Mechanical Treatment of a Rotating Magnetic Field Bloch Equations . . Solution of the Bloch Equations for Low HI Spin Echoes . Quantum Mechanical Treatment of the Spin Echo . Relationship Between Transient and Steady-State Response of a System and of the Real and Imaginary Parts of the Susceptipility . 2.12 Atomic Theory of Absorption and Dispersion ........•.•..
51 59
3. Magnetic Dipolar Broadening of Rigid Lattices . . . . . . . . •. . . . 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .• • . • . . 3.2 Basic Interaction . . . . . . . . . . . . . . . . . . . . . . . . •. . •. . . . . . . 3.3 Method of Moments ......................• . . • . . • . . . 3.4 Example of the Use of Second Moments .......•.•......
65 65 66 71 80
4. Magnetic Interactions of Nuclei with Electrons . . . . . . . . • . . • . • • . . 4.1 Introduction ...............................• • • • . • • . . 4.2 Experimental Facts About Chemical Shifts ......•..•..... 4.3 Quenching of Orbital Motion 4.4 Fonnal Theory of Chemical Shifts 4.5 Computation of Current Density ......•.............. 4.6 Electron Spin Interaction ............•..........•..•... 4.7 Knight Shift ..........................•..• ,.........
87 87 88 89 92 96 108 113
2.
Basic 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 2.10 2.11
1
1 2 4 11 II 13
17
20 25 29 33
35 39
46
VI VII
4.8 4.9
Single Crystal Spectra Second-Order Spin Effects-Indirect Nuclear Coupling
.
S. Spin-Lattice Relaxalion and Motional Narrowing of Resonance Lines . 5.1 Introduction . 5.2 Relaxation of a System Described by a Spin Temperature . 5.3 Relaxation of Nuclei in a Metal . 5.4 Density Matrix-General Equations . 5.5 The Rotating Coordinate Transfonnation .....•..•..•..... 5.6 Spin Echoes Using the Density Matrix . . 5.7 The Response to a 6-Function 5.8 The Response to a 1f/2 Pulse: Fourier Transfonn NMR . 5.9 The Density Matrix of a Two-Level System . 5.10 Density Matrix - An Introductory Example . 5.11 Bloch-Wangsness-Redfield Theory .............•........ 5.12 Example of Redfield Theory . 5.13 Effect of Applied Ahemaling Fields . 6. Spin 6.1 6.2 6.3 6.4 6.5 6.6
6.7 6.8 6.9 6.10 7.
Temperature in Magnetism liInd in Magnetic Resonance Introduction _ _ _ . A Predicrion from the Bloch Equations _ _.. The Concept of Spin Temperature in the Laboratory Frame in the Absence of Ahemating Magnetic Fields . Adiabatic and Sudden Changes . Magnetic Resonance and Saturation . Redfield Theory Neglecting Lattice Coupling . 6.6.1 Adiabatic Demagnetization in the Rotating Frame 6.6.2 Sudden Pulsing . The Approach to Equilibrium for Weak HI . Conditions for Validity of the Redfield Hypothesis . Spin-Lattice Effects _.....................•... Spin Locking, Tl". and Slow Motion ................•...
Double Resonance _ . 7.1 What Is Double Resonance and Why Do It? . 7.2 Basic Elements of the Overhauser-Pound Family of Double Resonance . 7.3 Energy Levels and Transitions of a Model System . 7.4 The Overhauser Effect ............................•... 7.5 The Overhauser Effect in Liquids: The Nuclear Overhauser Effect . 7.6 Polarization by Forbidden Transitions: The Solid Effect . 7.7 Electron-Nuclear Double Resonance (ENDOR) .
127 131
145 145 146 151 157 165 169 174
179 186 190 199
206 215
7.8 7.9 7.10 7:11 7.12 7.13 7.14 7.15 7.16 7.17 7.18 7.19 7.20 7.21 7.22 7.23 7.24
219 219 220
7.25 7.26
221 223
7.27 7.28
Bloembergen's Three-Level Maser ......••.••••.....•... The Problem of Sensitivity ................•....•...... Cross-Relaxation Double Resonance ................•... The B1oembe'1len-Somkin E
269 270 271 275 277 279 283 287
289 293
295 296 303 311
319 324
325 331 344 350 357
231 234
235 237 239 241
242 244
247 247
248 250 254
257 264
266
8. Advanced Concepts in Pulsed Magnetic Resonance 8.1 Introduction . 8.2 The Carr-Purcell Sequence . 8.3 The Phase Alternation and Meiboom-Gill Methods . 8.4 Refocusing Dipolar Coupling . 8.5 Solid Echoes . 8.6 The Jeener-Broekaert Sequence for Creating Dipolar Order 8.7 The Magic Angle in the Rotating FrameThe Lee-Goldburg Experiment . 8.8 Magic Echoes . 8.9 Magic Angle Spinning . 8.10 The Relation of Spin-Flip Narrowing to Motional Narrowing 8.11 The Fonnal Description of Spin·Flip Narrowing . 8.12 Observation of the Spin-Flip Narrowing . 8.13 Real Pulses and Sequences .......................•.... 8.13.1 Avoiding a z-Axis Rotation ..............•.•.... 8.13.2 Nonideality of Pulses . 8.14 Analysis of and More Uses for Pulse Sequence .....••....
367 367 367
369 371 371 380
384 388 392
406 409
416 421 421 422 423
VIII
IX
9.
Multiple Quantum Coherence . 9.1 Introduction _ . 9.2 The Feasibility of Generating Multiple Quantum CoherenceFrequency Selective Pumping . 9.3 Nonselective Excitation _ . 9.3.1 The Need for Nonselective Excitation . 9.3.2 Generating Multiple Quantum Coherence . 9.3.3 Evolution, Mixing, and Detection of Multiple Quantum Coherence . 9.3.4 Three or More Spins . 9.3.5 Selecting the Signal of a Particular Order of Coherence 9.4 High Orders of Coherence . 9.4.1 Generating a Desired Order of Coherence . 9.4.2 Mixing to Detect High Orders of Coherence .
10. Electric Quadrupole Effects
...... .......................... 10.1 Introduction 10.2 Quadrupole Hamiltonian - Pan J •••......••••.•......••• 10.3 Clebsch-Gordan Coefficients, Irreducible Tensor Operators, and the Wigner-Eckart Theorem
43I 43I 434 444 444
445
449 455
463 470 480
References
,
.............•.•.• , .••..•..•..•..•..•....
597 601 605
616 623 629
. 639
Author Index
...............................................
647
Subjed Index
...............................................
651
485
485 486 489 494 497
500
11, Electron Spin Resonance .,. , . , .. , , . , .. , . , .... , .... , ... , .. , . 11,1 Introduction , ,............................. 11.2 Example of Spin-Orbit Coupling and Crystalline Fields ., .. , 11.3 Hyperfine Structure 11.4 Electron Spin Echoes , _. . . . . 11.5 V", Center .................•....•..•................
503 503 505 516 524 533
1.2. Summary
555
Problems
557
Appendixes . A. A Theorem About Exponential Operators ............••..... B. Some Further Expressions for the Susceptibility ........•. , ... C. Derivation of the Correlation Function for a Field That Jumps Randomly Between ±ho , . D. A Theorem from Perturbation Theory ...............••..... E. The High Temperature Approximation ................• _.... F. The Effects of Changing the Precession FrequencyUsing NMR to Study Rate Phenomena .
579 579 580
x
Selected Bibliography
47I
10.4 Quadrupole Hamiltonian - Part 2 10.5 Examples at Strong and Weak Magnetic Fields 10.6 Computation of Field Gradienls ,............
.. . . . . . . . ••. . . . • . . • . • ••. . . .•. ••. . . . . . .•. .•. . . . .
G. Diffusion in an Inhomogeneous Magnetic Field , . H. The Equivalence of Three Quantum Mechanics Problems _ . !- Powder Patterns _ . _ . J. lime-Dependent Hamiltonians K. Correction Tenns in Average HamiltOnian TheoryThe Magnus Expansion _•••....
584
585 589
592 XI
1. Elements of Resonance
1.1 Introduction Magnetic resonance is a phenomenon found in magnetic systems mat possess both magnetic momenls and angular momentum. As we shall see, the term resonance implies thai we are in tune with a natural frequency of the magnetic system, in
this case corresponding to the frequency of gyroscopic precession of the magnetic moment in an external static magnetic field. Because of the analogy between the
characteristic frequencies of atomic Spectra, and because the magnetic resonance frequencies faJl typically in the radio frequency region (for nuclear spins) or microwave frequency (for electron spins), we often use the tenns radio frequency or microwave spectroscopy. The advantage of the resonance methoo. is that it enables one to select out of the total magnetic susceplibility, a particular contribution of interest - one that may, for example, be relatively very weak. The most spectacular example is, no doubt, the observation of the feeble nuclear paramagnetism of iron against a background of the electronic ferromagnetism. Resonance also pennits the gathering of precise, highly detailed magnetic infonnation of a type not obtainable in other ways. One of the reasons for the impact of magnetic resonance on physics is its ability to give infonnation about processes at the atomic level. In this book we seek to give some of the background necessary or useful to the application of magnetic resonance to the study of solids. Most of the book will be concemed with nuclear resonance, but the final chapters will focus on cenain problems particularly imponant for electron spin resonance. Many of the principles developed in the earlier ponions are, of course, equally applicable to nuclear or electron magnetic resonance. Our object is nOl to tell how to apply magnetic resonance to the study of solids. However, the activity in magnetic resonance has proceeded at such a vigorous pace, pouring out so many new concepts and results, that an author or lecturer faces an enonnous task in the selection of material. In this book, we shall use the study of solids as a son of ultimate goal that will help to delineate the topics for discussion and from which we shall attempt to draw most of the concrete examples of the more formal techniques. As we remarked above, we are concerned with magnetic systems that possess angular momentum. As examples, we have electron spins, or the nuclei of atoms. A system such as a nucleus may consist of many panicles coupled to-
gether so that in any given state, the nucleus possesses a total magnetic moment p. and a total angular momentum J. In (act the two vectors may be taken as parallel. so that we can write
The eigenvalues of this Hamiltonian are simple. being only multiples (7hHo) of the eigenvalues of I z • Therefore the allowed energies are
(1.1)
They are illustrated in Fig. 1.1 for the case I = 3f2. as is the case for the nuclei of Na or Cu. The levels are equally spaced, the distance between adjacent ones being 7hHo.
_E=-7hHom
where 7 is a scalar called the "gyromagnetic ratio". For any given state of a nucleus, knowledge of the wave function would in principle enable us to compute both p. and J. Hence we should find that the quantity 7 would vary with the state. Such calculations are beyond the scope of this book. Of course. in the quantum theory. p. and J are treated as (vector) operators. The meaning of the concept of two operators being "parallel" is found by considering the matrix elements of the operators. Suppose we define a dimensionless angular momentum operator I by the equation:
J= hI
.
(1.6)
'"
-1/2 1/2 3/2
Fig. 1.1. Energy levels or(I.6)
One should hope to be able to detect the presence of such a set of energy levels by some form of spectral absorption. What is needed is to have an interaction that can cause transitions between levels_ To satisfy the conservation of energy, the interaction must be time dependent and of such an angular frequency w that
fl
then has eigenvalues I(l + I) where I is either integer or half-integer. Any component of I (for example I z ) commutes with fl. so that we may specify simultaneously eigenvalues of both fl and I z • Let us call the eigenvalues I(l + I) and m. respectively. Of course m may be any of the 2I+1 values I. I -I •...• -I. The meaning of (1.1) is then that
Iiw = i1E
(1.3)
where 1J.l;' and Il;' are components of the operators I-" and I along the (arbitrary) x'-direction. The validity of this equation is based on the Wigner-Eckan theorem, which we shall discuss in Chapter to. We shall. for the remainder of this chapter, give a very brief introduction to some of the basic facts of magnetic resonance. introducing most of the major concepts or questions that we shall explore in later chapters.
.
-3/2
(1.2)
•
m=I.I-I •... ,-I
,
(1.7)
where 11E is the energy difference between the initial and final nuclear Zeeman energies. Moreover, the interaction must have a nonvanishing matrix element joining the initial and final states. The coupling most commonly used to produce magnetic resonances is an alternating magnetic field applied perpendicular to the static field. If we write the alternating field in tenos of an amplitude H~. we get a perturbing leno in the Hamiltonian of (1.8)
1.2 Simple Resonance Theory We shall wish. in later chapters. to consider both quantum mechanical and classical descriptions of magnetic resonance. The classical viewpoint is particularly helpful in discussing dynamic or transient effects. For an introduction to resonance phenomena, however. we consider a simple quantum mechanical description. The applicalion of a magnetic field H produces an interaction energy of the nucleus of amount -p.. H. We have. Iherefore, a very simple Hamiltonian:
1<=
-,,·H .
(1.4)
Taking the field 10 be Ho along Ihe z-direction. we find 1{ =
2
--yhHolz
(1.5)
The operator I z has matrix elements between states m and m'. (m'IIl;lm). which vanish unless m' = m±1. Consequently the allowed transitions are between levels adjacent in energy. giving
1iw = 11E = 7'lHo w = "tHo
or
(1.9) (1.9a)
Note that Planck's constant has disappeared from the resonance equation. This fact suggests that the result is closely related to a classical picture. We shall see, in fact. that a classical description also gives (1.9a). By studying the two formulations (classical and quantum mechanical). one gains a great deal of added insight. From (1.9a) we can compute the frequency needed to observe a resonance if we know the properties that detennine 7. Although such calculations are of 3
basic interest in the theory of nuclear structures. they would take us rather far afield. However, a simple classical picture will enable us to make a correct order-of-magnitude estimate of ..,. Let us compute the magnetic moment and angular momentum of a particle of mass m and charge e moving in a circular path of radius r with period T. The angular momentum is men
J = mvr = m
21rr 2
T
(1.10)
while the magnetic moment (treating the system as a current loop of area A carrying current i) is ~:iA
(1.11)
FIg. 1.2. Energy levels
e 'lfr 2
-(1.12) c T Comparison of the expressions for 1J and J therefore gives us "'/ = e/2mc. Besides enabling us to make an order of magnitude estimate of the expected size of "'/. for our purposes the imporlant resuh of this fonnula is that large masses have low ",/'s. We expect about a factor of 1.000 lower.., for nuclei than for electrons. In fact. for magnetic fields of 3,()(X) to 10,000 Gauss. electronic systems have a resonance at w/21r = 10,000 MHz (the 3 em microwave region), whereas nuclear systems are typically 10 MHz (a radio frequency). Of course one can always change w by changing H o, but in most cases it is advantageous to use as large a magnetic field as possible, since the quanta absorbed are then larger and me resonance is correspondingly stronger. In later sections. we shall comment somewhat more on typical experimental arrangements.
rOT
1
=;
yAHo +1(2
N.
~+ :
N_WH _(+) - N+W(+)_H
(1.13)
Without as yet attempting to compute W(+) ..... (_} or W(_) ..... (+). we note a famous formula from time-dependent perturbation theory for the probability per second Po; ..... " that an interaction Vet) induces a transition from a stale (a) with energy Eo to a Slate (b) whose energy is E,,:
p._, 2; l(blVla)I'.(E. _ E, _ =
Since i = (elc)(llT), we get ~.
-1(2 _ _-,-,,---_ N_
r",,)
(1.14)
:::: l(bIVla)1 2, we note that po; ..... " is the same as the rate P" ..... o. Such' an argument describes many situations and leads 10 the condition W(+)_H: WH_(+) " W.
Since
l(alVlb)1 2
dN+ = W(N_ _ N+) dt
(1.15)
It is convenient to introduce the variable n:::: N+ - N_. the difference in population of the twO levels. The two variables N+ and N_ may be replaced by n and N. using the equations (1.16) (1.160)
Substitution of (1.16a) inlo (1.15) gives us
1.3 Absorption of Energy and Spin-Lattice Relaxation We now wish to go a step further to consider what happens if we have a macroscopic sample in which we observe a resonance. For simplicity we consider a system whose nuclei possess spin ~ (Fig. 1.2). Since there are many nuclei in our macroscopic sample. we shall specify the number in the two m states and - , by N+ and N_. respectively. The total number of spins N is a constant, but application of an ailernating field will cause N+ or N_ to change as a result of the transitions induced. Let us denote the probability per second of inducing the transition of a spin with m = to a state m:::: -~ by W(+)-+(_). We shall denote the reverse transition by W(-l ..... (+} . We can then write a differential equation for the change of the popu ation N+.
+l
+!
4
dn -::::-2Wn dt the solution of which is n:::: n(0)e-2Wt
(1.17)
(1.18)
where nCO) is Ihe value of n at t =O. We note mat if inilially we have a populalion difference, it will eventually disappear under Ihe action of the induced ttansitions. The rate of absorption of energy dEldt is given by computing the number of spins per second that go from the lower energy to the upper. and by subtracting the number that drop down. emitting energy in the process: (1.19)
5
Therefore, for a net absorption of energy, n must be nonuro; that is. there must be a population difference. We see that when the upper state is more highly populated than the lower, the net absorption of energy is negative - the system supplies more energy than it receives. This state of affairs is the basis of the os· cillators or amplifiers known as masers (microwave amplification by stimulated emission of radiation) or lasers (for light amplification). We see that if the equations we have put down wen:: complete. the resonant absorption of energy would eventually stop and the resonance would disappear. A more serious difficulty is seen if we assume W = 0 (that is. we do not apply the alternating magnetic field). Under these circumstances our equations say that dN+/dt = O. The populations cannot change. On the other hand. if we applied a static field to a piece of unmagnetized material. we should expect it to become magnetized. 11Je preferential alignment of the nuclear moments parallel to the field corTCsponds to N+ being greater than N_. (N_ = 0 would represent perfect polarization, a state we should not expect to find at temperatures above absolute zero). The process of magnetization of an unmagnetized sample. therefore, requires a net number of transitions from the upper to the lower energy state. In the process. the spins give up energy - there is. so to speak. a heat transfer. Therefore there must be some other system to accept the energy. If we ask how big a population difference will eventually be found, the answer must depend upon the willingness of the other system to continue accepting energy. Speaking in thennodynamic terms, the heat flow will continue until the relative populations N_/N+ correspond to the temperature T of the reservoir to which the energy is given. The final equilibrium populations N~ and N~ are then given by
N~
NO
=e-AE/kT =e-'Ylllo/kT
We must postulate. therefore. that there exists a mechanism for inducing transitions between N+ and N_, which arises because of the coupling of the spins to some other system. Let us denote the probability per second that such a coupling will induce a spin transition upward in energy (from + _ -) by Wr, and the reverse process by W 1. 1ben we have a rate equation
.
(1.21)
Let us again introduce the variables Nand n; but now we no longer can assume equality of the two transition probabilities. since we know such an assumption would not give the preference for downward transitions which is necessary for the establishment of the magnetization. In fact. since in the steady-state dN+/dt is zero. (1.21) tells us that
N~
Wr N~ = W! 6
(1.22a)
W! =e'Ylllo/kT
.wr
It is natural to wonder why the argument given to show the equality of W(+) _ (-) and W(_)_(+) does not also apply here. The ~Iution of this paradox is th~t the thennal transition requires not only a couphng but also another system 10 an energy state that permits a transition. We can illustrate by assuming that the reservoir has only two levels whose spacing is equal to that of the nuclear system. If the nucleus and reservoir are initially in the states of Fig. 1.3a given by the crosses conservation of energy is satisfied by simultaneous transitions indicated by the :m.ows. The nucleus may therefore give up energy to the lattice. On the other hand, if both systems are in the upper state (Fig. 1.3b), the simultan~~us transition cannot occur because it does not conserve energy. The rate of translUon of the nucleus will therefore depend not only on the matrix elements but also on the probability that the reservoir will be in a state that permits the transition.
,. aT Reservoir
Nucleus
2~
b
Reservoir
Nucleus
"
(al
(bl
FIg. 1.3. (a) A possible ~ransition. (b) A forbidden tnmsition
(1.20)
+
dN+ -;u=+N-W!-N+Wr
By using (1.20). we find that the ratio of W 1 to Wr is not unity but rather is
Thus if we label the nuclear stateS I and 2 with populations N 1 and N2. and label ~he lattice states (a) and (b) with populations N a and Nb. the number of transitions per second. such as shown in Fig. 1.3a, will be (1.23)
where W lb _2a is the probability per second of such a. tn:"sition u~r the condition that the nucleus is actually in state 1 and the lattice IS actually 10 state (b). The steady-state condition is found by equating the rate of such transitions to the rate of the inverse transition: N1NbWlb_2a = N2Na W 2a _ 1b
(1.24)
Since the quantum theory requires that Wlb_2a = W 2a _ 1o • we see that in thermal equilibrium,
(1.22)
7
That is, the nuclear levels will have the same relative populations as do those of the lattice. The nuclear population will therefore be in thermal equilibrium with that of the lattice. Note. moreover, that for this simple model, we can compute
WI and W": (1.26)
so that Wl and W! arc seen to be unequal. We now leave our special model and return to (1.21). By making the substitutions of (1.16a) for N+ and N_, we find
dn
dt = N(W" -
W)) - neW" + WI)
(1.27)
which can be rewritten as
dn
no-n
----dt Tl
where
Wl- WI)
no =N ( W!+Wl
(1.28) I
TI
=(WI+W))
(1.29)
Since the solution of (1.28) is
n = no + Ae- t / 7i
(1.30)
(where A is a constant of integration), we see that n~ represents the thermal equilibrium population difference, and TI is a characteristic time associated with the approach to thennal equilibrium. T I is called the "spin-lattice relaxation time". For example, if we deal with a sample that is initially unmagnetized, the magnetization process is described by an exponential rise to the equilibrium: n=no(l_e- t / T1 )
(1.31)
That is, TI characterizes the time needed to magnetize an unmagnetized sample. We may now combine the two rate equations for dn/dt to find the combined transition rate due to both thennal processes and transitions induced by the applied alternating field:
dn no - n -=-2Wn+--dt T1
(1.32)
In the steady state. (1.32) tells us that
no
n = 71-:+-;2"W~T'-1
(1.33)
dE
W
dt = nliwW = n o1iw 1 +2WTt
(1.34)
We shall see later that W is proportional to the square of the alternating magnetic field. Therefore (1.34) tells us that we can increase the power absorbed by the nuclei by increasing the amplitude of the alternating field. as long as 2WT1
Gorter's. We have also seen that the rate of absorption is related to the transition rate W. An estimate of the size of the resonance absorption is basic to a decision about whether or not a resonance might be observed. We shall wish to consider ho'w to calculate W. Moreover, since no resonance line is perfectly sharp, we expect that the factors governing the width of [he spectral line will be of interest. Closely related is the question of what magnetic field to use in the relation w = '"'tHo, for the nuclei are never bare. There will be magnetic fields due to electrons as well as due to other nuclei. which must be added to the external field. These fields produce effects of greatest interest. such as the splitting of the proton resonance of ethyl alcohol (CH3CH20H) into three lines of relative intensities 3:2: I. They are also responsible for the fact that there is a nuclear resonance in ferromagnets even in the absence of an applied static magnetic field.
Therefore, as long as 2WT,
9
2. Basic Theory
2.1 Motion of Isolated Spins - Classical 'freatment We begin our sludy of (he basic theory with a classical description of the motion of a spin in an external magnetic field H, assuming Ihal H may possibly vary with lime. H will produce a IOrque on the magnelic moment 11. of amount Il x H. If we applied a magnetic field to an ordinary bar magnet, mounted with bearings so thai it could turn at will, the magnet would attempt to line up along the direction of H. If H were constant in time and if the bearings were frictionless. the magnet would actually oscillale about the equilibrium direction. If the bearings were nOI frictionless, the oscillalions would die out as the magnet gave up energy to the bearings. until eventually it would be lined up along H. When the magnet also possesses angular momentum, the situation is modified. since it now acts like a gyroscope. As we shall see, in the event of frictionless bearings, the moment would remain al fixed angle with respect to H (providing H is constant in time), but would precess about it. The conversion of energy back and forth between potential energy and kinelic energy would nOI occur. It would still be lrUe, however, thai if Ihe bearings possessed. friction, the magnet would eventually become parallel to a static field H. As we shall see, the friction corresponds to relaxation processes such as TI. The equation of motion of the magnet is found by equating the torque with the rale of change of angular momentum J.
dJ
dt
= I' x H
(2.1)
Since p. = "(J, we may eliminate J, getting (2.2)
This equation, which holds regardless of whether or not H is lime dependent, tells us that at any instant the changes in p. are perpendicular to both p. and H. Refer to Fig.2.1 and consider the tail of the vector p. as fixed; the tip of the vector is therefore moving out of the paper. The angle 8 between p. and H does not change. If H is independent of time, the vector p. therefore generates a cone. One can proceed with the solution of (2.2) by standard methods of differential equations for various assumed time dependences of H. We shall find. it 11
Fig. :1.1. Relation of /' to Jl
most useful for our future work, however, to introduce a special technique: the use of a rotating coordinate system. Consider a vector function of time F(t), which we may write in terms of its componenls F:t(t). Fy(t). Fz(t), along a set of rectangular coordinates. In terms of the corresponding unit vectors i, i, and k, we have (2.3)
Ordinarily we think of i. i, and k as being constant in time. but we shall wish to be more general. Since their lengths are fixed, they can at most rotate. We shall assume they rotate with an instantaneous angular velocity {}. Then
di
dt =
n
.
x ,
= I' x (,H + n)
n
,
(2.9)
most~eneral
+I "L
(2.10)
time-dependent solution !li(t) is therefore cm u I,m e-(i/h)E",1
(2.11)
m=-1
where the em's are complex constants. We may compute the expectation value of any observable by means of !P(t). as we can illustrate with the x-component of magnetic moment: (2.12)
(2.6)
(2.7)
Equation (2.7) tells us that the motion of p. in the rotating coordinate system obeys the same equation as in the laboratory system, provided we replace the actual magnetic field H by an effective field He:
He=H+-
The
wet) =
where we have introduced the symbol 6FMt. representing the time rate of change of Fwith respect to the coordinate system i, i, k. For example, when 6F16t = 0, the components of F along i. i, and k do nOI change in time. By making use of (2.5), we can rewrite the equation of motion of JL in terms of a coordinate system rotating with an as yet arbitrary angular velocity {}:
'I' lit
Em = -"(hHom
WI ,m (t) = UI ,m e-(i/h)E",t
(2.5)
"
'I' 6't+{}xJL=p.x"(H or
We have seen that the quantum mechanical description of a spin in a static field gave energies in terms of the quanlum number m, which was an eigenvalue of the component of spin 1z parallel to the static field Ho. The energies Em were
The corresponding eigenfunctions of the time-independent SchrOdinger equation may then be denoted by u/,m' The time-dependent solution corresponding to a particular value of m is therefore
dF . dF:t di .dFv dj dFz dk dt = '"""dt + F:t dt + J"""dt + F v dt + k dt + F z dt
,F =-+{}xF
2.2 Quantum Mechanical Description of Spin in a Static Field
(2.4)
The time derivative of Fis therefore
.dF:I: .dFv dFz • = .d -+ ,-+ k-+ n x (.F: +,'F, +kF) tdtdt l:VZ
We can now readily solve for the morion of p. in a static field H = kHo by choosing {} such that He = 0. That is, we lake {} = -"(Hok. Since in this reference frame 6p./6t = 0, p. remains fixed with respect to i. i. and k. The motion with respect to the laboralory is therefore that of a vector fixed in a set of axes which themselves rotate at n = -"(Hok. In other words, JL rotates at an angular velocity n = -,,(Hok with respect to the laboratory. The angular frequency"(Ho is called the "Larmor frequency". We are struck by the fact that the classical precession frequency {} is identical in magnitude with the angular frequency needed for magnetic resonance absorption, as found by elementary quantum theory. Let us therefore look more closely at the quantum mechanical description.
We have emphasized that the expectation value of !J:t. (!J:t) will vary in time by explicitly writing it as a function of time.
1 We write a variable of integralion dr in lhe expression for the expedation value, in analogy to that which we would do for a spatial coordinate :1:, V, z or angular coordinates 6, .p. For spin, the nolation is to be thought of as a symbolic representation of the scalar product of the two funClions >it(L) and 11rl/i(L)
(2.8) 13
12
/
By using the fact that P~ :: 'rhI~, and that !V(t) is given by (2.11) we find (p~(t»::
L:
'rliC:n,cm(m'll",lm)e(i/h)(Em,-Em)1
(2.13)
m,m'
In order to gain funher insight into tile physical significance of the general expression for (,ux(t», (2.13), we now consider the fonn it takes for a spin of By using the fact Ihat the diagonal matrix elements of I", vanish, we get
!.
(,u",(t)) "" 'rll[ ci'/2c-1/2(! II", I -
where
(m'II",lm)
==
J
uimd",ulmdr
l1 t . 1/ 2Clj2 ('II + c_ -'2' ~ I''2' )ei"( o l
(2.14)
is a time-independent matrix element. Expressions similar to (2.13) would hold for any operator. We denole that the expectation value will in general be time dependent, will consist of a number of tenns oscillating hannonically and that the possible frequencies '
t)e -i"(l1ot (2.19)
It is convenient to define a quantity wo '" 'rHo. As we have seen, wo is the angular frequency we must apply to produce resonance and is also the classical precession frequency. By utilizing the fact that (!Il",lis the complex conjugate of IIzl!), and using the symbol "Re" for "take the real part oC', we get
t)
(-!
Em/-Em
are just those which correspond to the frequency of absorption or emission be. tween states m and m'. Of course it was the assumption that observable properties of any quan.tum sys~em had 10 be given by expressions such as (2.13), which was .the baSIS of Heisenberg and Born's fonnulation of the quantum theory in matnx fonn. Since matrix elements (m'II~lm) vanish unless m' :: m± I, we see that all ~e tenos of (2.13) ha:e .an angular frequency of either +-rHo or -'rHo. ~lf sum. m~st also contam Just 'rHo. The expectation value (p",(t» therefore OSCillates In tIme at the classical precession frequency. It is convenient at this point to introduce the famous raising and lowering operators I+ and I-, defined by the equations r+:: I", +iIy
I-:: I", - iIy
(2.16)
We may express I", or I y in tenns of I+ and'r by solving (2.16), getting I + _ 1.=,([ +1)
I I V =2i(p-r)
(2.17)
The operalors are called "raising" or "lowering" because of the effect they produce when they Operate on a function 'Ill ,m: r+UI,m::
VI(I + I) - m(m + 1)UI,m+1
I-UI,m '" VI(I + I) - m(m -l)ltl,m_1
(2.18)
I+ turns Ul,m into a ~unction whose m value has been raised by one unit. We see,. Iherefore, tha,t (m II+lm) vanishes unless m' = m + I, while (m'II-lm) vams~es unless m = m - 1. Van Vleck [2.1) has characterized these as "sharper" se.lec n~n rules than those of the operators Ix or ly, which may join a state Ul WIt h eIther uI,m+l or UI,m_l. ,m
14
(2.20)
~I~
h
We evaluate the matrix element by means of (2.17) and (2.18), getting
(W.I- ~)=~.
It is convenient at this point to express the c's in tenns of two real, positive quantities a and b. and two other real quanlities (which may be positive or negative) a and p: Ct/2 "" ae
ia
,
i C_lj2 = be /1
(2.21)
The nonnalization of the wave function gives us a 2 + b2 "'" 1. These give us
(1'",(0) "'" 'rhab cos (a -
fJ + wot)
(2.22a)
Similarly we find
(Py{t) = -'rhab sin (a -
p + wot)
(p,(t» = 1h(a' - 6')/2
.
(2.22b)
We note that both (1'",) and (lJy) oscillate in time at the Lannor frequency 'rHo, but that (P:) is independent of time. Moreover the maximum amplitudes of (1'",) and (Py) are the same. If we define
(,,) " i(p.) + j(p,) + k(p,)
(2.23)
and utilize the fact that (p",)2 + (py)2 "'" constant, a fact readily verified from (2.23), we see that (I-') behaves as does a veClOr making a fixed angle with. the z·direction, precessing in the x-y plane. In teons of polar coordinates 6, t/J (see Fig.2.2), any vector A may be written as
A",=Asin 6 cos t/J Ay=Asin6sint/J A:=Acos6 .
(2.24)
15
."1g.2.2. Relationship of the components A"" A,o and A, of a vcctor A to the polar angles 6, 1/>, and the magnitudc A
:
A,
o
A
By means of algebraic manipulation one can show that (p~) =
"'(It .
2""
Sill ()
cos ¢
h
(J1.y) = "'(2 sin () sin ¢ (J1.z) =
,Ii
2"" cos
()
For our example: (2.25)
¢ = {3 - 0: - wot 2 l+cos(} 2
~~
One may look on (2.26) as a formal change of variables, of course, but the results of (2.25) tell us that there is a simple physical significance; the expectation value of the operator It acts as does a vector of length "'(nfl, whose direction is given by the spherical coordinates (), ¢. If the orientalion is specified at any time, it can be found at future times by recognizing that it precesses at angular velocity wo in the negative rP direction. The orientation may be specified quite arbitrarily (by specifying a or band (3 - a). We emphasize that an arbitrary orientation can be specified, since sometimes the belief is erroneously held that spins may only be found pointing either parallel or antiparallel to the quantizing field. One of the beauties of the quantum theory is that it contains features of both discreteness and continuity. In terms of the two quantum states with m = ± ~ we can describe an expectation value of magnetization which may go all the way from parallel to antiparallel, including all values in between. Thus a wave function with a = b has an expectation value corresponding to a magnetization lying somewhere in the x-y plane (that is, with vanishing z-component). Just where in the plane it points is given by the complex phase a - {3, as well as the time at which we wish to know the orientation. 16
Pl/21/2=a2 Pl/2-1/2 =
provided
a=
It is useful to consider briefly what we should expect for the wave function if we took a sample of many noninteracting spins which were in thermal equilibrium. There will be a wave function for each spin, but in general it will not be in one of the eigenstates (m = +~ or m = -~); rather it will be in some linear combination. For a given spin, there will be a particular set of values for a, b, 0', {3. The values will differ from spin to spin. For example, we have a distribution of the quantity 0' - {3 that gives the spin orientation in the x-y plane at t = O. If the spins are in thermal equilibrium, the expectation value of the total magnetization must be parallel to the magnetic field. We expect, therefore, that there will be no preference for anyone value of 0' - {3 over any other. That is, the spins will have a random distribution of 0' - (3. On the other hand, since the spins will be polarized to some extent, we expect to find a larger than b more often than b is larger than a. That is, the average value of a must be larger than the average value of b. Since an observable quantity can be expressed in the fonn of (2.13), we see that we can specify either the individual em's or the complex products c:n,c m , which we shall label Pmm/ for convenience.
P_I/2_1/2=b2
abe i (Q'-,8)
P- 1/ 21 /2 = abei(f.l-Q')
We may consider the Pmm,'s to be the elements of a complex matrix P. Notice that the diagonal elements (m = m') give the probabilities of occupation of the various states, while the off-diagonal elements are closely related to the components of magnetic moment perpendicular to the static field. We shall make use in a subsequent section of the average of the matrix P over a statistical ensemble. The statement that in thennal equilibrium the magnetization will be parallel to the field amounts to saying that (he average over the ensemble of Pmm , for m l '" m is zero, whereas the average for m = m' is the Boltzmann factor describing the probability of finding the state occupied. (Of course, in the quantum theory, even for a number of spins with identical wave functions, any experiment that COUIllS the number of spins in the various m states will find a statistical distribution not related, however, to temperature.)
2,3 Equations of Motion of the Expectation Value TIle close correspondence of the classical and quantum mechanical treatments is made particularly clear by examination of a differential equation relating the lime variations of the expectation values (Pz), (Jly), and (J1.z). The equation is based on a well-known fonnula whose derivation we sketch. 17
Suppose we have a pair of wave functions !li(t) and w(t), both of which are solutions of the same SchrOdinger equation:
_":.aop i
at
= HOP
(2.27)
Let us have some operator F that has no explicit time dependence. Then
~ JP* F!lidr:::
*J
w*(hF - Fh)!lidT
i
f,[H,!,] -1'Hoi[Iz. I z ] -yHoly
(2.34.)
(2.28)
Similarly,
(2.29)
dIy -::: --yHolz dt dlz ::: O . (2.34b) dt These equations are the component equations of the vector operator equation
This equation is readily derived from the fact that
into which we substitute expressions for the time derivative taken from (2.27).2 It is convenient to write (2.28) in operator form. There is no problem with the right-hand side: It is simply (i/h)(hF - Fh). For the left-hand side we must define some new notation. We define the operator dF/dt by the equation (2.30) That is to say, dF/dt does !lot mean to take the derivative of F with respect to t. Such a derivative vanishes, since F does not contain the variable t. Rather dF/dt is a symbol that has the meaning of (2.30). By using dF/dt in this symbolic sense, we have (2.31) where [h, F] is the usual commutator HF - F'1l. We may use this fonnalism to compute the time derivative of the expectation values of /lz. /ly, and j1.z. We define the X-, y-, z-axes as being fixed in space but with the z-axis coinciding at an instant with the direction of the magnetic field. (In this way we include both static and time-varying fields.) Then (2.32) We shall wish to use the commutation relations for the components of angular momentum, all of which may be obtained by cyclic permutation from (2.33) Then
2 To p~ove (2.28), one must use the fad that F is an Ile~mitian operalor. (See discussion in Sed. 2.5).
18
dI, = ..dt ::: :::
dI -::: Ix -yH where
(2.35)
dI dt
(2.36)
dt
.dlz
.dly
kd1z
:::1di+ 1 di+ 'dt
Therefore, since Jl- ::: -yliI, we have the equation for the expectadon value of magnetization,
(,,>
de,,> = x 1H (2.37) dt which is just the classical equation. In words, (2.37) tells us that the expectation value of the magnetic moment obeys the classical equation of motion. Equation (2.37) was derived for the expectation value of a magnetic moment of a single spin. If we have a group of spins with moments Jl-k' for the kth. spin, their total magnetic moment Jl- is defined as
,,= L>k
(2.38)
k
If the spins do not interact with one another, it is easy to prove that (2.37) also holds true for the expectation value of the total magnetization. Since, in practice, we measure the results of a number of spins simultaneously, the experimental measurements of magnetization measure the expectation value of the various components of magnetization. That is, the experimentally detennined bulk magnetization is simply the expectation value of the total magnetic moment. Therefore the classical equation correctly describes the dynamics of the magnetization, provided the spins may be thought of as not interacting with one another. It is important to bear in mind that (2.37) holds true for a time-dependent H, !lOt simply a static one. Therefore it enables us to use a classical picture for studying the effects produced by alternating magnetic fields. We turn to that in the next section. 19
2.4 Effect of Alternating Magnetic Fields The effect of an alternating magnetic field H~(t) = H%o cos wt is most readily analyzed by breaking it into two rotating compon~nts, each. of amplitude HI. one rotating clockwise and the other counterclockwise (see Fig. 2.3).
The time dependence of HI can be eliminated by using a coordinate system that rotates about the z-direction at frequency w,.. In such a coordinate system. H1 will be static. Since the axis of rotation coincides with the direction of Ho. Ho will also be static. Let us take the :t-axis in the rotating frame along HI. Then (2.41) becomes
i:
y
II,
x F1g. 2.3. Oeeomposition or a linear oecillating field into two rotating e1el1'lenls
~x
[k(w,. + "1Ho) + i"1Hd
(2.42a)
Notice that we have encountered two effects in making the transformation of (2.41) to (2.42a). The first is associated with the derivative of the rotating unit vectors and gives the term W,. The second is associated with expressing the vectors Ho and H, in tenns of their components in the rotating system and gives rise to the conversion of HI from a rotating to a static field. Equation (2.42a) may be rewriuen to emphasize that near resonance w,. +"1HO ~ 0, by setting w,. = -w. where w is now positive (we assume here that "1 is positive). Then
=" x 7[ (Ho - ~)k+H,il
We denote the rotating fields by H R and HL:
HR =H1(i cos wt + j sin wt) H L = H1(i cos wt - ; sin wt)
=
=~xHeff
(2.39)
NOie that HL and HR differ simply by a replacement of w by -w. Since one component will roGue in the same sense as the precession of the moment and the
other in the opposite sense. one can sho'" that near resonance the countcrrotating component may be neglected. We shall 1......_ thai approximation in what follows. Alternatively we can assume Ihal we are finding the exact solution of a problem in which the experimcnlal arrangement has produced a rotating field; for example, by use of two identical coils at right angles to each other and with alternating currents 90 degrees out of phase. We shall assume we have only the field H R, but this is no loss in generality because the use of a negative w will conven it to HL. In order to reserve the symbol w for a positive quantity, we shall introduce the symbol Wz, the component of w along the z-axis. w,. may therefore be positive or negative. We may, therefore, write (2.40)
(2.42b)
where Heff=
k(Ho -~) +H)i
Physically (2.42b) stales that in the rotating frame. the moment aclS as though it experienced effectively a static magnetic field Heff. The moment therefore precesses in a cone of fixed angle about the direction of Heff at angular frequency -yHeff- The situation is illustrated in Fig.2.4 for a magnetic momer.! which, at t = O. was oriented along the z-direction.
,
,
H., y
y
which will give us either sense of rotalion, depending on the sign of w,.. We now ask for the equation of motion of a spin including the effects both of HI (t) and of the static field Ho = kHo.
x (oj
(2.41)
20
(b)
FIg. 2.4. (a) Eff«.tive field. (b) Motion or the moment Jl in the rotating coordinate system
21
We notice that the motion of the moment is periodic. If it is initially oriented along the z·direction, it periodically returns to that direction. As it increases its angle with the %-direction, its magnetic potential energy in the laboratory reference system changes (in the laboratory system the magnetic energy with respect to Ho is much larger than that with respect to HI, so we customarily neglect the laller). However, all the energy it takes 10 tilt JL away from HO is returned in a complete cycle of JL around the cone. There is no net absorption of energy from the alternating field but rather alternalely receiving and returning of energy. Note thal if No is above resonance (thai is, Ho >wh), the effective field has a positive %-component, but when Ho lies below that resonance (Ho <w!"(), the effective field has a negalive %-cornponent. If the resonance condition is fulfilled exactly (w = "'IHo), the effective field is then simply ill I. A magnetic moment that is parallel to the static field initially will then precess in the y-z plane. That is, it will precess but remaining always perpendicular to H,. Periodically it will be lined up opposed to Ho. If we were to tum on HI for a short time (Ihat is, apply a wave train of duration t w ), the moment would precess through an angle 8 = ..,Hlt w . If t w were chosen such that 8 = r, the pulse would simply inven the moment. Such a pulse is referred to in the literature as a "180 degree pulse". If 8 = 1fn. (90 degree pulse), the magnetic moment is turned from the %-direction to the y-direction. Following the turn-off of HI, the moment would then remain at rest in the rotating frame, and hence precess in the laboratory, pointing nonnal to the static field. These remarks suggest a very simple method of observing magnetic resonance, illustrated in Fig. 2.5. We put a sample of material we wish 10 study in a coil, the axis of which is oriented perpendicular to Flo. In thennal equilibrium there will be an excess of moments pointing along Flo. Application of an alternating voltage to the coil produces an alternating magnetic field perpendicular to Ho. By properly adjusting Ht and t w ' we may apply a 90 degree pulse. Fol-
II.
II,
,, Ii , f
"
,.)
I I
-
II,
- - -
-
\
-
,
-+
_.1
the magnetization M will tum with Ho. always remaining aligned along Ho as Flo turns. To prove this theorem, let us assume w to be a constant in the z-direction. We can take it perpendicular 10 110, since a component parallel to [-[0 produces no erfect. The relationships are shown in Fig. 2.6 at t = 0, with M and Ho taken parallel to each other and pointing in the X -direction in the laboratory. If we choose a reference frame x, y, %rotating at angular velocity nR = w, Ho appears static. but we must add an effective field nRh. Choosing the z- and Z-axes as parallel, and x to coincide with X at t = 0, the effective fields and magnetization at t = 0 are shown in Fig. 2.7. The effective field in the rotating frame is static and given by
,
-- -
z
\
w
'e)
.'1g.1.5. (a> Coil cOlilainin.g sample. 'n thcnnal equilibrium all cxceu or moments is parallel to 110. (b) and (c:) Followmg il 9O-degree pulse, t~ excess momellls precess perpendicular to Jlo
22
..,Ho>w
nR W Hclf=Ho+-=Flo+-
\ (b)
lowing the pulse, the excess magnetization will be perpendicular to Ho and will precess at angular frequency "'IHo. As a result, the moments will produce a flux through the coil which will alternate as the spins precess. The resultant induced emf JIlay be observed. What we have suggested so far would indicate that the induced emf would persist indefinitely, but in practice. the interactions of the spins with their surroundings cause a decay. The decay may last in liquids for many milliseconds, but in solids it is more typically 100,IS. Even during that shan time, however, there are many precession periods. The technique we have described of observing the "free induction decay" (that is, decay "free" of HI) is a commonly used technique for observing resonances. It has the great virtue of enabling one to study the resonance signal in the absence of the voltages needed to produce HI. Since oscillators always generate noise. such a scheme may be advantageous. One interesting application of the rotating rererence frame is to prove the following theorem, which is the basis of another technique for producing resonance signals. Suppose we have a magnetic field [-[0 of fixed magnitude whose direction we may vary (no other magnetic field is present). Let the magnetization M be parallel to Ho at t = O. We may describe the changing direction of Flo by an angular velocity w. Then the theorem states that if
,
y I
23
-------
Fig. 1.7. Magneti\l:Rtion M Rnd e{fedive field H eff in the rotaling coordinate system :e,y,z. The magnetizalion will precess aboul the effective field in the cone or angle (J shown
Fig. 1.8. (al Magnetization M lind effective field Heff in the rotating frame, with M parallel to lleff. (b) The situation exactly at resonance, having approached resonllnce slowly, with A
w
lIo -T'
II"
M M
II, )'
l,)
•
II,
lb)
2.5 Exponential Operators It will be useful to consider the quantum mechanical equivalent of the rolating coordinate transformation, but to do so, we shall need to employ several useful relations. We review them here for the convenience of the reader. Suppose we have two wave functions, tP and t/i. that satisfy appropriate boundary conditions and have other satisfactory properties for some region of space. and suppose we have an operator F. F may be, for example, a component of spin. 1be operator is said to be Hennitian when
x
M will precess about fIerr, making an angle 8 such that
(2.44)
w
tan 8:: (2.43) 7 HO M will therefore remain within an angle 28 of fIo. We see that if wl-rHo <: I, M and fI 0 remain parallel. The fact that the magnetizalion follows the direction of the magnetic field when the field changes direclion sufficiently slowly is described by the term adiabatic. By uliliz.ing this principle. one can tum to the case of a rotating magnetic field HI of frequency w, perpendicular to a static field Ho. If one starts far below resonance, the magnetiz.ation is nearly parallel to the effective field in the rotating
JH?
frame + [(w/-y) - HoJ2. As one approaches resonance, both magnitude and direction of the effective field change. but if resonance is approached sufficiently slowly. M will remain parallel to fIeff in the rotating frame according to the theorem we have just proved. Thus. exactly at resonance. the magnetization will lie along HI> making a 90 degree angle with Ho (Fig.2.8). If one were to continue on through the resonance, the magnetization would end up by pointing in the negative z-direCl"ion. This technique of inverting Mis very useful experimentally and is called "adiabatic inversion".
24
where the integrals are over the region of space designated. To prove that an operator is Hermitian requires some statement about the conditions lJi and tP are to satisfy, as well as a definition of the region. For example. if F is an operator involving derivatives. the proof that it is Hermitian may involve transforming the volume integ-,al to a surface integral and requiring the integrand of the surface integral to vanish on the surface of the region. Hermitian operators are important because their expectation values and eigenvalues are real. lltercfore any operator that corresponds to a physically observable quantity must be Hermitian. Thus Ihe operators I:e' I y • and I: are Hermitian. If they are Hermitian, it is easy to show from (2.44) that the operators I+:: I:e + iIy and I- :: I;~ - iIy are not. In the theory of functions. it is useful 10 define the exponential function of the complex variable z: z2
z3
e: :: I + Z + - + - + 21 3! the power series converging for all z. We define the function F2 p3 e F :: 1 +p+-+- + 2! 3!
25
similarly, where F is now an operator. We shall be particularly interested in the function e
iF
.
(iF)2
(iF)3
= I +IF+~+3!+ ...
= exp (iwotI:)!li(O)
(2.45)
By using the series expansion, one can show that if F is Hennitian, exp (iF) is nol. In fact
J(e iFtJi)*lJidr = J p*e- iF !lidr
!li(t) = exp [ - (i/h)( -")"IiHoI:)t]w(D)
(2.46)
The exponential function of operators obeys some of the same algebra as does the function of ordinary number, but as usual with operators, care must be taken whenever two noncommuting operators are encountered. Thus, if A and E are two operators, one can verify by means of the series expansion that
where Wo :: ")"Ho. We know that H 0 produces a rotation of the magnetic moment at angular velocity a given by n = -")"Hok. We shall call such a rotation "negative", since the component of angular velocity along the z-axis is negative. It is log· ical to suppose, then, that !li(t) must correspond to the function w(D), referred, however, to axes rotated in the negative direction through an angle wot. Thus exp(-iI:4')w(O) should correspond to a function identical to w(O) referred to axes rotated through the positive angle 4'. If we compute the expectation value or matrix elements of, for example, Ix., we find
J
w*(t)Ix.w(t)dr = =
(2.47.) only if A and B commute. Likewise,
=
(2.47b) only if A and E commute. If A and B do not commute, another useful equation may still hold. Let us define C as the commutator of A and B:
[A,BJ=AB-BA=C. Suppose that C commutes with both A and B:
[A,C]=O
[B,C]=O
Theo e(A+B) = eAeBe-Cj2 = eC/2eBeA
This theorem is proved in Appendix A. Use of the exponential function provides a particularly simple method for obtaining a fonnal solution of SchrOdinger's equation if the Hamiltonian does not depend explicitly on time. That is, if !li(t) is the solution of
Ii a~(t)
-Tat = 1ilJi(t)
(2.48)
then we can express !li(t) in tenns of its value at t = 0, l1i(D), by the equation
~(t) = e-(i/')"'~(O)
(2.49)
Equation (2.49) may be verified by direct substitution into (2.48). tf, for example, we consider the motion of a spin in a magnetic field so that 1i;:: -")" 1tHoIz , 26
(2.50)
J J J~·(O)J.,(t)~(O)dT
[eiwotl·i[i(O)j*I:r.ei""Otl·!li(O)dr
w*(O)e-i""OII. Ize iwotl• !li(O)dr
(2.50.)
where
(2.50b) The last line defines the operator I z " We can give a simple interpretation of (2.50) as follows: The first integral, which gives (Iz(t»), corresponds to a precessing angular momentum arising from the effect on a time-independent operalor I z of a time-dependent funclion !P(t). The last integral describes the effect on a timedependent operator Iz'(O of a wave function p(D), which is independent of time. Since the precession is in the negative sense, the first integral involves a fixed operator and a wave function fixed with respect to axes that rotate in the negative sense. Therefore the last integral must describe an operator rotating in the positive sense with respect to the "fixed" wave function !li(D). It is a simple matter to show that I z ' is related to I z through a rotation of axes. Let us consider
(2.51) We wish to find /(4'), to see what meaning we can ascribe to it. Of course we could simply expand the exponentials and, using the commutation laws, try to reduce the function to something tractable. A simpler method is to show first that f(4') satisfies a simple differential equation and then solve the equation. We have
df ;:: e-i1.t/l(-iIzIz + iIzIz)eil.t/l
d1
(2.52)
But, since [Iz,Izl = iIy , 27
df
:=
e-iIEtP IyeilEtP.
(2.53)
&1 Likewise
(2.54)
2,6 Quantum Mechanical Treatment of a Rotating Magnetic Field We shall now use the exponential operators to perfonn the quantum mechanical equivalent of the classical "rotating coordinate" transfonnation. We shall consider a magnetic field HI, which rotates at angular velocity Wz, in addition to the static field kHo. The total field H(t) is then
Therefore H(t):= iHI cos wzt + JB I sin wzt + kHo
f(¢):= A cos ¢+B sin ¢
where we must evaluate the constants of integrations. (As we shall see, the "constants" are actually operators.) Clearly, A '" f(O), but from (2.51), f(O):= I~. Likewise, B:= t(O):= I y , using (2.53). In this way we get I~I
(2.57)
and the SchrOdinger equation
(2.58)
== e- il • tP I~eil.tP:= I~ cos ¢+ ly sin ¢
By using (2.55) of the preceding section, we can write the Hamiltonian of (2.58) (2.55)
as (2.59)
Izl
== e-iIE
The quantities I~" Iyl' and Izi are clearly the components of angular momentum along a set of axes Xl, y', z' rotated with respect to x, y, z, as shown in Fig. 2.9. Therefore we see that we can use the exponential operator exp (il z ¢) to generate rmations. It is frequently useful to work with the raising and lowering operators I+ and I-. Then the first two equations of (2.55) can be rewritten as e-il.tPI+eil.tf; :=I+e-itf;
and
(2.56a) (2.56b)
We are tempted to try to "remove" the operator exp (iwztIz ) from I", and transfer it onto 1Ji, much as the reverse of the steps of (2.50) of the preceding secrion. Accordingly we let (2.60)
The physical interpretation of (2.60) is that IJi and 1Ji' differ by a rotation of axes through an angle wzt (a rotating coordinate transfonnation). Then, using (2.60)
8w . • t1,alJi' -:=-iw Ie- .Iw • '1 EIJiI+e- 1W _
at
"
zz
at
(2.61)
We may substitute (2.60) and (2.61) into (2.58), multiply both sides from the left by cxp (iwztI z ), and obtain
x'
•
x
Fig.2.9. Relation of axes "', Y to ",', y' and thO) angle tP
h alJi'
-Tat:= -[hew: +,Ho)Iz +,hHIIz]1lI
(2.62)
In (2.62) the time dependence of HI (t) has been eliminated. In fact we recognize it as representing the coupling of the spins with an effective static field
k(Ho+~)+iHI 28 29
the effective field of our classical equations. The spins are therefore quantized along the effective field in the rotating coordinate system, the energy spacing being "'{fiHeff' The wave function .pI given by (2.60) is related to the function tJi by a coordinate rotation, the "forward" motion of Iz; relative to a stationary l/J having been replaced by a stationary I z and "backward" rotating tJi'. As usual, resonance occurs when Wz ::;:::: - ",(Ho. If we define the transfonned Hamiltonian 11.' by
1t =
-[(Ttw z + "'{ IlH o)Iz + "'{II HI I:a:]
(2.63)
we can formally solve (2.62):
tJi'(t)
=e-(i!i)'H'IW'(O)
(2.64.)
whence, using (2.60). !li(t)::: e- iw,I1'e-(i!i)1t'I!Ji'(O)
=
(2.64b)
=
{Note that at t O. !li(0) lli'(O)j. Equation (2.64b) gives us a particularly compact way to express the solution of SchrOdinger's equation when a TOtating field is present. We can illustrate the use of the wave function of (2.64b) by computing the time dependence of the expectation value of Jlz. Of course we know already what the resuh must be. since we have proved that the classical picture applies. Let us for simplicilY assume that HI is applied exactly at resonance. Then, from (2.63),
1t = -...,IlH lIz
(2.65)
Then we have. using (2.64b) and (2.65),
(p,(l»
=
If we define
J[e-i"",
If the- magnetization lies along the z·axis at t = 0 so that {J.ly(O» = 0, we get (2.71) Thus the z-magnetizalioo oscillates in time at ...,H 10 corresponding to the precession of {po) about HI in the TOtating reference frame. It is important to note lhat in lhis picture. which neglects all interactions of spins with one another or the lattice. the magnetization continues oscillating between +{1l:(0» and -{J.l:(0» indefinitely. This behavior is very different from that which we should expect from a time-independent transition probability such as we assumed in Chapter 1. The time-independeOl transitions occur only if some physical process spoils the coherent precession about H t in the rotaling reference frame. Another approach to solving (2.62) enables us to demonstrate a very interesting and fundamental property of spin particles, their so-called spinor nature. Formally, this property describes what happens to a wave function under a rotation. One can discuss the problem by fonnal mathematical melhods. as in group theory (see below). Here. however. we display the mathematical result by physically generating a rotation with an alternating magnetic field tuned exactly to resonance so that (Ho + w:h) = 0 in (2.62). Let Utj2 and U_Ij2 be the eigenslales of the spin operator I z • They are independent of time. Since they form a complete set for a spin ~ particle, we can express any function, such as I/J', as a linear combination of Ulj2 and U_lj2 with coefficients a and b. If I/J' is itself a function of time, ,p'(O, the coefficients a and b must also be functions of time:
!
(2.72)
tf• e i "l'll1 I.. ttP'(O>t I z [e -i"",LI. ei"l'Il\/"ttJi(O)JdT.
(2.73)
WI.
Muhiplying by U;j2 from Ihe left, integrating over spin space, and utilizing the (2.67)
and lise the fact thul I z and I z are Hermitian, we get
J.p'
(O)e -i""l tlz Izei""l If.0 tJi(O)dT
f.ct that (11[.1~) = (il da
="'(Ii Jv*(O)e -i""l t1 rei"", LI. Ize -i"".t1 ze i""l tI" !P(O)dT
= "'(Ii
!II.I-1)
=0, we get
1
T dt = 1IiHtb(~I[.I-~) (2.68)
7
I
(2.69)
(2.74)
Similarly we get
Ii db
By using (2.55), we can write
30
(2.70)
(2.66)
WI ="'(H]
{,tz(t»
{Jlz(t» = -{,JII(O» sin wit + {Jtz(O» cos wit
,p'(0 ::: a(t)u Ij2 + b(Ou_Ij2
J~'(l)I,,~(l)dT
= "'(It
Substituting in (2.68) we get
-
dt
= 1"Ht a ( -
!II.lt)
Utilizing the fact that {!II:,.,j -!)::: i and {two simultaneous differential equations to find
illzl!) =!
we can solve these 31
+ ib(O) sin (WI t(2.) b(t) :::: ia(O) sin (WI t(2.) + b(O) cos (WI tn)
aCt) "" a(O) cos (WI tn.)
(2.75)
(al
{bl -
(2.76) is the classical precession frequency about HI in the rotating frame. The spinor properly is revealed by considering a 211'" pulse. We recall that such a pulse causes (he expectalion value of the magnetization vector to undergo a rotation about HI which returns it to its initial value. Denoting the pulse duration as t2 .... we have. then Wlt2 ... :::: 211'"
6('2.) = -6(0)
Bark
«
I
p;f
y'
y'
·1
".
~~
y'
. ".
'wisled strip
{dl
X·
6(••) = 6(0)
The t ..isled slrip ....itll its ends joiI'led
(2.78)
These relationships show that after a 2x- rotation the wave function has nbl relurned to its original value but has instead changed sign. For the wave funclion 10 return to its initial value. the pulse length, t..... must lead to a 4x- rotalion. Thus. if WI t w '" 4x-
(2.79)
The general property that a 211'" rotalion produces a sign reversal of ¢ and thai a 4x- rolation is needed 10 gel t/J back to its initial value is referred to as the "spinor" properly of ",. It is shared by wave functions associated with spins of ~. ~, etc. The existence of spinors is well known in group theory. For example. in solid-state physics the property is referred to with the tenn a crystal double group [2.2]. The wave functions of particles wilh spins of 0, 1,2, elC. return to their original value under rotations of 2'11". There is somcthing unselliing aboUl finding that a 2'11" rotation does not return one to one's starting point! It is perhaps comforting in this connection 10 think of a Mobius strip, which is shown in Fig. 2.10. As is explained in the picture, one must go around the strip twice to reach the starting point. Thus. we have a physical manifestation or representation of (2.73). Referring to (2.22) we sec that though the wave function changes sign on a 2"'11" rotation, the expectation v.lllIes of the spin components [;e, [y, I: do not. The question arises whether or not the spinor nature is physically observable. The answer is yes. The first explicit demonstration was done by two groups: by Rauch et al. [2.3], and independently by Werner et al. [2.4]. Methods of providing a test had been proposed earlier by Bernstein {l.5] and by Aharmwv and Susskind [2.6]. The essence of the idea is to produce a spatial separation of the spin-up part of the wave function from the spin-down part so Ihal one can act on the two parts independently. Following a spatial region in which a magnetic field acts on only one component of the wave function, the two components are recombined (0
32
"
(2.77)
which gives. using (2.75).
!.
x'
",
y
where
a(t.... ):::: a(O)
x front
"
Fig.2.IO. A Mobius strip can be enviuged by starting with a strip of pllper ... hose fronl (a) "nd back (b) sides can be di$linguished by minting them black and ....hite respectively. The paper is then twisted (e) and the ends joined (d). Suppose one then starts on the bh.ck !lUnOKe at the point Pa. moves on thu surface to i15 other end (point Qo) llt which point one crosses over to point f\v on the white surface. Ancr going the length of the ....hile surface, one lIrrives at point Qw, adjACent to the original slarling point. Po. Thus, one has been around the strip twice to reach the starting point
study their interference. Thus, if one leaves the spin-up ponion alone. it provides a fixed phase reference for (he spin-down function. When one subjects the spindown function to 211" and 4'11" rOtations by passing it through a region of magnetic field, one finds that the interference intensity is the same for 0 and 4Jf rotations, but different for a 21f rotation. Ingenious NMR experiments demonstrating the spinor propeny have also been perfonned by Stoll and co-workers [2.7-2.9).
2.7 Bloch Equations Both quantum mechanical and classical descriptions of the motion of noninteracting spins have in common a periodic motion of the magnetization in the rotating frame. For example, if ')'Ho:::: wand if the magnetization is parallel to Ihe static field at t "" 0, Ihe magnetization precesses around HI in (he rotating frame, becoming alternately parallel and ami parallel to the direction of the static field. Viewed from the laboratory frame. the magnetization is continuously changing its orientation with respect to the large static field. However, Ihe energy that 33
must be supplied to tum the spins from parallel to antiparallel to the static field is recovered as the spins return to being parallel 10 the static field. Thus there is no cumulative absorption over long times but rather an alternate absorption and recovery. The situation is reminiscent of whal we described in the first chapter prior to introduction of the coupling to the thermal reservoir. (We note that there the system, however, simply equalized populations, whereas our present model predicts an alternating reversal of populations. The two models must therefore be based on differing assumptions.) Without contact to a reservoir, we have no mechanism for the establishment of the magnetization. By analogy to the equation
dn
dt
=
no-n
T1
(2.80)
and recognizing that M: = ,,/lin/2, we expect that it would be reasonable for M: to be established according to the equation
dM:
--= dt
Mo -M: TI
1 - - =" T2= (2.81)
where Mo is the thennal equilibrium magnelization. In terms of the static magnetic susceptibility Xo and the static magnetic field Ho, we have
Mo = xoHo
(2.82)
We combine (2.81) with the equation for the driving of M by the torque to get
dM" di""" =
M, - M. (M H) TI + "/ x :
(2.83)
Funhermore we wish to express the fact that in thermal equilibrium under a static field, the magnetization will wish to be parallel 10 Ho. That is, the x- and y-components must have a tendency to vanish. Thus (2.84)
dM M dt Y = "'(M x If)y - T: We have here introduced the same relaxation time T2 for the x- and y-directions, but have implied that it is different from Tl. That the transverse rate of decay may differ from the longitudinal is reasonable if we recall that, in contrast to the longitudinal decay, the transverse decay conserves energy in the static field. Therefore there is no necessity for transfer of energy to a reservoir for the transverse decay. (This statement is not strictly true and gives rise to important effects when saturating resonances in solids, as has been described by Redfield. We describe Redfield's theory of saturation in Chapter 6, beginning with Sect. 6.5). On the other hand, the postulate of the panicular (exponential) form of relaxation we have assumed must be viewed as being rather arbitrary. It provides
34
a most useful postulate to describe certain important effects, but must not be taken too literally. According to (2.84), under the influence of a static field the transverse components would decay with a simple exponential. (This result is readily seen by transfomling to a frame rotating at ,,/Ho, where the effective field vanishes.) A possible simple mechanism for T2 for a solid in which each nucleus has nearby neighbors arises from the spread in precession rates produced by the magnetic field that one nucleus produces at anDlher. If the nearest neighbor distance is 1', we expect a typical nucleus to experience a local field Hloc"'" Il/r J (due to the neighbors) either aiding or opposing the static field. As a result, if all nuclei were precessing in phase at t = 0, they would get out of step. In a time T such that "'(HlocT ::! I, there would be significant dephasing, and the vector sum of the moments would have thus diminished significantly. Since T must therefore be comparable to T2, a rough estimate for T2 on this model is "'(Hl oc
(2.85)
,,/2ft
often about 100 JLS for nuclei. Equations (2.83,84) were first proposed by Felix Bloch and are commonly referred to as the "Bloch equations". Although they have some limitations, they have nevenheless played a most imponant role in understanding resonance phenomena, since they provide a very simple way of introducing relaxation effects.
2.8 soluiion of the Bloch Equations for Low HI At this stage we shall be interested in the sollllion of the Bloch equations for low values of the alternating field, values low enough to avoid saturation. We immediately transform to the coordinme frame rotating al w: taking HI along the x-axis and denoting H o + (w:h) by 11 0 . Then
dM: _ dt
~1 H
- - --"'(I' Y
1+
Mo ~ M: T[
(2.86,)
(2.86b)
(2.86c)
Since M~ and My must vanish as HI -10, we realize from (2.86a) that in a steady Slate, !vI: differs from Mo to order H~. We therefore replace M: by Mo in (2.86c). The solution is funher facilitated by introducing M+ = M~ + iM y . By adding (2.86b) 10 i rimes (2.86c), we get 35
Mx(t) ==
(2.87)
(x' cos wi + i' sin wt)Hxo
(2.93)
defining the quantities X' and Xii. By using (2.90) and (2.93), we get I
,.
(2.88)
a=-+-y10I
T,
- I
Xo 2
......
X = -WO.L2
Therefore
M+ = Ae-o·t + i-yMoHI
(2.89)
1/T2 + i,ho
If we neglect the transient term and subsulUte Mo = xoHo. and define Wo :: 7Ho. = -w, we get
Wz
(2.90)
II
X =
It is convenient to regard both Mx(t) and Hx(t) as being the real parts of complex functions M~(t) and H~(t). Then, defining the complex susceptibility X by I
Mil "" XO(WO T 2) 1 + (W -Wo )'T' H) 2
(2.93,)
Xo...... I -w,p, 2 1+(w-wo)2Ti
.
X=X -IX
I
(wo - W)T2 1 +(w - wo)2Ti
/I
(2.94)
and writing
Equations (2.90) show thai the magnetization is a constant in the TOtating reference frame. and therefore is TOtating at frequency w in the laboratory. In a typical experimental arrangement we observe the magnetization by studying the emf it induces in a fixed coil in the laboratory. If the coil is oriented with its axis along the X -direction in the laboratory, we can calculate the emf from Icnowledge of the time-dependent component of magnetization
MX along the
(2.95)
Hi(i) = H XOeiWl
we find
M~(t) = xHi(t)
or
MX(t) == Re HxH xoeil.Jt)}
(2.%) (2.%,)
X -direction. y
x
"'
Fig.2.1l. Rotating
axes~,
II relative to lll.borillory llXetl
X, Y
Although (2.92) and (2.96a) were arrived at by considering the Bloch equations, they are in fact quite general. Any resonance is characterized by a complex susceptibility expressing the linear relationship between magnetization and applied field. Ordinarily, if acoil of inductance Lo is filled with a material of susceptibility xo, the inductance is increased to Lo{l + 47fxo), since the flux is increased by the factor I + 47fXo for the same current. In a similar manner the complex susceptibility produces a flux change. The flux is changed not only in magnitude but also in phase. By means of (2.93-96), it is easy to show that the inductance at frequency w is modified to a new value L, given by (2.97)
By referring to Fig. 2.1 I, we can relate the laboratory component Mx to the components M~ and My in the rOlating frame. Thus
Mx = M z cos wt + My sin wt
(2.91)
where X(w) = i(w)-ixl/(w). It is customary in electric circuits to use the symbol j for yCT. However, in order to avoid the confusion of using two symbols for the same quantity, we use only i.3
If we write the magnetic field as being a linear field, Hx(t) = Hxo cos wi
2HI = H xo
(2.92)
then we see that both M z and My are proponional to H xo, and we can write
3 In practice, the sample never complC!lely fills all apace, and we must introduce the ~fil1illg factor~ q. Ita calculation del>ends 011 a knowledge of the spatial variation of the
alternating field. Then (2.97) bccomCl!l L = Loll
+ "... qX(loI)J 37
Denoting the coil resistance in the absence of a sample as Ro, the coil impedance Z becomes
Z :::: iLow(1 + 41Tx' - i41TX") + Ro :::: iLow(l + 41TX/) + L ow41TX" + Ro
(2.105)
(2.98)
The real part of the susceptibility Xl therefore changes the inductance, whereas the imaginary part, XII, modifies the resistance. The fractional change in resistance IJ.RI Ro is
Low. -L1R = - 1TX " Ro
Ro
:::: • 1rX"Q
Utilizing the fact that H xo :::: 2H I this can be rewriuen as
(2.99)
whe~ v = W/21T. Thus, if one knows the Q, the volume of the coil, and the power available from one's oscillator, one can calculate how strong an H xo (or HI = Hxo/2) one can achieve, frequently a very useful quantity to know. In (2.105) the units are [H 1 ]: Gauss; [?]: erg/s; [v]: Hz; [V]: cm 3. Using 1w = 107 erg/s, an alternative expression in mixed units is HI =
lOPQ
(2.106)
vV
where we have introduced the so-called quality factor Q, typically in a range of 50 to 100 for radio frequency coils or 1,000 to 10,000 for microwave cavities. Assuming unifonn magnetic fields occupying a volume V, the peak stored magnetic energy produced by an alternating current, whose peak value is io, is
with [Hd: Gauss; [P]: W; [II]: MHz; [V]: cm J . The particular functions X' and XII, which are solutions of the Bloch equations, are frequently encountered. They are shown in the graph of Fig. 2.12. The tenn Lorentzjan line is often applied to them.
(2.100) The average power dissipated in the nuclei P is
p:::: !iijIJ.R:::: ~iijLow41ri'
(2.101)
By substituting from (2.100), we find
I H'xoX "V P :::: ~w
(2.102)
This equation provides a simple connection ,between the power absorbed, X", and the strength of the alternating field. We shall use it as the basis of a calculation of X" from atomic considerations, since the power absorbed can be computed in tenns of such quantities as transition probabilities. Since X' and XII are always related, as we shall see shortly, a calculation of XII will enable us to compute Xl. Moreover, we recognize that the validity of (2.102) does not depend on the assumption of the Bloch equations. Another useful fonnula can be obtained from (2.100) and (2.101), relating the average power dissipated in the coil resistance, Pc, to the strength of Hxo. Since Pc is half the peak power dissipmcd in the coil, (2.103)
Fig. 2.12. X' and X" rrom the Bloch equations ploued versus z "" (wo - ",,)T2
At this time we should point out that we have computed the magnetization produced in the X -direction by an alternating field applied in the X -direction. Since the magnetization vector rotates about the Z-direction, we see that there will also be magnetization in the Y -direction. To describe such a situtarion, we may consider X to be a lensor, such that C(t)M Ct' -
XCl"Cl'
H (toe i""t
0:
= XYZ , ,
o:l=X,Y,Z
In general we shall be interested in Xxx.
2.9 Spin Echoes
Solving for (~)iij and substituting into (2.100) we get
Hl ov : : PcQ 8,
38
w
(2.104)
Just after finishing graduate studies, Erwin Hahn burst on the world of science with his remarkable discovery, spin echoes [2.10]. His discovery provided the
39
key impetus to the development of pulse methods in NMR. and must therefore be ranked among the most significant contributions to magnetic resonance. What are spin echoes and why are they so remarkable? Suppose one applies a 1fn. pulse to a group of spins to observe the free induction signal which follows turn-off of the pulse. According to the Bloch equations. the free induction signal decays exponentially with a time constant T2' For solids. T2 is a fraction of a millisecond, corresponding to line widths of several Gauss. For liquids, the line widths are typically much narrower. corresponding perhaps to times of several seconds. Such lines are a good deal narrower than the usual magnet homogeneity. As a result, the inhomogeneity-induced spread in precession frequency causes the spins in one portion of the sample to get out of phase with those in other JXlnions. The free induction signal arises from the sum total of all portions of the sample. As the separate portions get out of step, the resultant signal decays. The decay lime is of the order of l/(-y.dH) where tj,H is the spread in static field over the sample. fJalm made the remarkable discovery that if he applied a second 7rn. pulse a time T afler the first pulse, miraculously there appeared another free induction signal at a time 21' after the initial pulse. He named the signal the "spin echo". To produce a signal at the time of the echo, the spins must somehow have gotten back in phase. The great mystery of the spin echo was what made the spins get back in phase again? Was the echo a challenge to basic concepts of irreversibility? Was there a Maxwell demon at work producing the refocusing? fJalm discovered spin echoes experimentally, but was soon able to derive their existence from the Bloch equations. This solution showed that as one varied T. the echo amplitude diminished exponentially with a time constant T2. Thus the echo provided a way of measuring line widths much narrower than the magnet inhomogeneity. Understanding the physical basis of echo formation has led to much deeper insight into resonance phenomena in general and pulse work in particular. The essential physical ideas of refocusing can be most easily seen by considering a pulse sequence in which the first pulse produces a rotation of 7I:n.. the second a rotation of 1f. Such a sequence we denote as a 7rn.-7r pulse sequence. It was invented by Carr [2.11] based on a vector model proposed by Purcell for the 1rn.-7f/2 echo [2.10]. and was first described in a famous paper by Carr and Purcell [2.12). Consider a group of spins initially in thennal equilibrium in a static magnetic field H in the z-direction. The thermal equilibrium magnetization Mo then lies along H as shown in Fig. 2.1301. We assume there is a spread in magnetic fields over the sample, and take the average value of field to be Ho. We first analyze what happens if we can neglect the effect of Tl and T2. We apply a rotating magnetic field HI at t = 0 with frequency w. tuned to resonance at the average field Ho. Thus
w=,Ho 40
(2.107)
,
, M,
}---y x
x
(b)
, M!
----
/
0
,
--" --
I"" 0'
, --,
--,,
.""7'----i-- y / --'
x
x (d)
I'" T+
, FIg. 2.13a-<. The formation of a spin echo by means of a 7t /2pulse sequence viewed in the rotating reCeren<:e frame. (I) At t = 0- the magnetizalion, Mo is in thermal equilibrium lying along lhe :-direction. (b) shows lhe magnelizalion immediately after lhe 71:/2 pulse. In (c) an element of magnetization, 6M, has precessed an utra angle' owing to the magnelic. field inhomogeneilY. (d) shows the effect of the 7t pulse on 6M. In (e) we !lee that at time 2T all elements of magnetization hilove refocused along the +y-direclion
7t
Y x (t')
r= 2T
By proper adjustment of the pulse length t p , we can generate a 'lrn. pulse. In our discussion we consider H} to be sufficiently strong that t p is negligibly short. We designate the time just before or just afler the initial pulse by 0- and 0+, respectively. It is convenient to view the behavior of the spins in the reference frame that rotates at,Ho. with the x-axis defined as lying along HI. Let the 7r/2 pulse rotate Mo to lie along the negative y-axis (Fig.2.13b). This sense of rotation corresponds to a spin which has negative 'Y, such as an electron. For a positive 'Y, the rotations are positive in the left-handed sense. We give the laller case as a homework: problem for the classical derivation of the echo, and treat it quantum mechanically in the next section. If there were not inhomogeneity in the static field, all spins within the sample would precess at 'YHo, so that the magnetization of every portion of the sample would remain 41
It/,
oriented along the -V-axis. The existence of inhomogeneity leads to a spread in precession rates and dephasing. Consider what happens during a time interval,. In any small region of the sample, the magnetization will remain in the x-V plane since we are neglecting T, processes. But at the end of" the direction of within that plane will advance from the -v-direction by some angle which we call 6, given by
pulsl:
11
l'ulsl:
oM
6='1oH,
where
6H=H-HO
'"
(2.108)
o
(2.109)
FIg.2..14. The deeay of the echo from I = 2T onward ill the same function of time as the decay of the free indudioll signal from j 0+ onwards. Note that the buildup before t = 2T is the mirror image in time of the decay aner 2T. Nole furl her that no free induction signal is produced immediately following the ... pulse
represents the inhomogeneity in H. The situation is shown in Fig.2.13c. (Note that oH may be either positive or negative, so that the "advance" may be either positive or negative.) Let us assume we can control the phase of the oscillating voltage in the second pulse so that the H t again lies along the +x-direction in the rotating frame. 4 Suppose now we apply a 'II" pulse at t =', again of negligible duration. We denote the time just before and just after the pulse by t = , - or t = respectively. The situation just after the'll" pulse is shown in Fig.2.13d. Noting the orientation of oM, we immediately see that during a second time interval " oM will again advance through the same angle 6, which will bring it exactly along the positive y-axis at t = 2T. The argument applies to all spins. no matter what oH they experience, because the result does not depend on the angle of advance. Though all the spins are in phase at t = 2" they get out of phase again owing to the field inhomogeneity, so the free induction signal decays. Note that its fonn as a function of time during the dephasing from t = 2, onward must be identical to the fonn of the decay following the initial 1fn pulse (see Fig. 2.14). The buildup of the echo signal just prior to t = 2r is the mirror image in rime of the decay after t = 2,. lt is easy to see now what will be the effect of the T , and T2 tenns of the Bloch equations. During the first time interval" the components of in the x-v plane will decay exponentially with T2' and a z-componem will
,+.
oM
4 With so-called coherent pulse apparatus, a highly stable oscilll\lor generl\tes l\ steady-state alternating voltage which is fed into a circuil which amplifies the rf voltllge only during the lime thalli gl\le voltage is applied, lhereby producing a strong rf signal coincident with the gale pulses. The output voltage is fed to the sample coil to generate lhe lf l . rr lhe slable oscillator is tuned exactly to ~nance,..., = ""0, the phasc of the Ifl is always tIle same in the ""0 reference frame. This method of operalioll, together wilh so-called ph~e coherent deledion of the free indudion decays, WI\S first introduced independently by Solomon [2.131 and by Spokm and Stichter [2.'",15). In 12.15]' there ill a dCS<:.ription of the rel\SOns for using phase-coherent detection. In a famous paper on nuclear relaxation in alkali metals, lheir colleagues, I/o/comb and Norberg 12.161 introduce lhe use of signaillveraging to improve the signal to noise ratio in pulse experiments, lind point out lhal coherent detection is necessary to achie\'c the full potential of II noise integration. 42
Helm
"'rl:l: ;",1"CI;(/II sigl/a/
oM
,
2,
=
develop exponentially with T,. The 'II" pulse will invert the z-component which in the :.:-y has developed, so that it does not contribute to the component of plane existing at t = ,+. During the next interval" the comp:ment in the :.:-y plane will continue to decay via the T2 tenns in the Bloch equations. As a result, the size of the magnetization producing the echo signal M(2T) wiU obey
oM oM
M(2T) '" Moe- 2T / T1
(2.110)
If the second pulse is a 1fn pulse, as in Hahn's original experiment, the pulse at t = , will put some of the z-component of into lhe :.:-y plane. One might
oM
then wonder whether or not the echo contained an exponential depending on T•. A delailed analysis shows that lhat does not happen. This result is most easily seen by pretending the magnetization developed along the z-axis arose from an independent set of spins which possess no net x-y magnetization. With respect to the new spins the situation just prior to lhe second pulse is like Fig.2.13a, with oM lying along the +z-direction. The 1f/2 pulse puts the magnetization into the :.:-v plane. The situation at t = 2T is identical to that of Fig.2.13c, in which all the spins have dephased. Thus the Tl induced magnetizalion does not affect the echo. To observe the signal, one would need to refocus - which requires still another pulse. Thus with three pulses one could see T I effects. HO/111 in fact found such echoes. He observed that if he applied a third pulse at time T (hence T-, after the second pulse), he produced an echo at 2(T-,) after the second pulse, or at t = 2T-, after the first pulse. He found, in addition, echoes at T + T, 2T - 2T, and 2T. The various echoes induced by a third pulse are commonly referred to as "stimulaled" echoes. In liquids, the diffusional motion pcnnits a nucleus to move between different parts of the sample where the precession rates may differ. As a result, during a spin echo the dephasing during the first interval, may differ from the rephasing during the second interval T, and the echo is diminished. The effect is of great practical utility as a way of measuring diffusion rates in liquids. This 43
was discovered by Hahn, and reported in his first spin echo publication [2.10]. Although Hahn gives correctly the expression for the effect of diffusion on the decay of the transverse magnetization following a single pulse. there is an error in his expressions for the effect of later pulses, corrected by Carr and Purcell [2.12] in a paper which also showed how the effect of diffusion could be eliminated (Sec!. 8.2). Carr and Purcell showed thai diffusion led to a decay of the echo peak magnetization AI, given by
M(2r) "" Moexp where
8z - 3 [-..,'1 (aH)'2DT3]
(2.111)
we have assumed
H - Ho.
(2.112)
as is the case for a magnetic field of axial symmetry about the z-axis. We derive this resuh in Appendix G. As a preliminary, in Appendix F we treat a simpler case in which bodily motion affects the struClure of a nuclear absorption spectrum. In an actual experiment, since HI is not infinite, the HI pulses have a nonzero duration. Let t./ 2 be Ihe duration of a 7(0. pulse if one were exactly at resonance (,H,t./ 2 '" 7(0.). We call this the "7(0. pulse" in what follows even though it is not exactly so for spins which are off resonance. We then ask what influence the nonzero duration of the pulse has on the time al which the echo occurs. Indeed, one might ask whether or not the echo is fonned perfectly under these circumstances. To discuss the problem, we define the times T" and ~ shown in Fig. 2.15. We define T as the time between the end of the 7fo. pulse and the start of the 7f pulse, T' as the time from the end of the 'If pulse to the peak of the echo (assuming one is fonned!).
1\ o
,
'"
,. -----<
=
6Hitt.O
"
'IIn-A9
':5~~=H~'~"]-"ii:':::]i~==-" 'r 6H x H, ,,
\
[one 01 Ihe precessing mignetiulion 6H
",,, along Hi> and the x'-axis along Herr. The magnetization contribution oM will precess about Herr. in a cone which makes an angle 1fo. - .69 with respect to %' where tan.69 '" SH/HI. With this as a starting point, one can then work through the details of the precession. If HI is much bigger than the line width, then .69 <: I, and one can reexamine the derivation given for the echo. We put the details of Ihe derivation as a homework problem. One finds, keeping only corrections linear in .d9, that the "7fo. pulse" is a 1fo. pulse for rotations about x'. It puts the magnetization at an angle .69 = oH/H, with respect to the negative y-axis. Thus, at the Slart of the 1[" pulse the spin now makes an angle .d9+"'(oHT with respect to the y-axis instead of SHT of (2.108). The 'Ir pulse can be shown to introduce no elTor in the orientation of SM in the x·y plane. Thus at a time T after the end of the 7f pulse when '
Echo Frn induction decOly
l-, _
f\ I Echo
I peak
Fig.2.tS. Defillition of ~he times .. lind ..' us~d to discuss echo for"'l\~ion when one considers the effect of lhe finit.c si1.e of //1, h~nce lhe nonzero length of lhe 1r/2 lind 11" pulses
To analyze this problem, let us suppose Ihat Ihe resonance is broadcned symmetrically by a field inhomogcncity and that HI is tuned exactly to the center of the resonance, W '" ,Bo. Consider an elcment of magnetization, oM, which is off resonance by H - Ho '" oH. Then the effective field in the rotating frame aCling on this magnetization is shown in Fig.2.16. We take the x-axis
44
,
\
z( ~~)
,
,.
Fig. 2.16. The effec~ of l\ finite Til on lhe precession of l\ component of magnetization, 6M, which is off resonance by a ",agnetic field 6//. //1 lies along the z-axis. The ",'-axis, which liCli along the effective field H~ff, makes an angle .d6 tan-I (6// / I1 t ) with the z-axis. Thus, iJ.M precesses about H~ff in a COlle of allgl~ ./2 - D6
(2.113) the magnetization lies along the +y-direction. This gives I
T''''T+-,HI
(2.114)
which is independent of 6fI. Thus, the oM's from all parts of the resonance line are collinear, and an echo is fonned at this lime. Equation (2.114) can also be written as
T''''T+~t1["/2
(2.115)
This expression was derived on the assumption that SH« HI. Numerical simulations for SH>HI' however, show that the correction tenn, IhH\, remains accumte to within 10% even for lines much broader than H, [2.17).
45
2.10 Quantum Mechanical Treatment of the Spin Echo
dN = NP(lto)dllo
The spin echo can also be derived quantum mechanically. Of course, since (2.37) shows that the expectation value obeys the classical equations of motion, one immediately knows that the classical derivation is true for the quantum mechanical expectalion value. However, there is added insight from following the development of the actual wave function in time. We consider,then, a sequence ofevems in time shown in Fig. 2. 17, which shows H I versus time. At t = O. HI of frequency w is turned on along the x-axis in the reference frame rotating at -kw. It remains on until time tit producing a 1(12 pulse. From tl to t2 the spins precess freely. A second pulse is applied from t2 to t3, where the interval is chosen to make this a 11" pulse. We denOle t2 - tl as T. Since we will be working solely in the rotating frame, we will omit from t/J and 1t the prime used earlier to designate quantities in the rotating frame.
H,
(2.119)
(N is the total number of spins). See Problem 2.3 for a classical study of the effecJs of this distribution function. We can simplify things by assuming that, while HI is on, HI ;:;. Ito for all spins, so that we can approximate 'H during the pulses by
(2.120.) In between pulses, when HI '" 0, we have (2.12Ob) Although 1{ is time dependent, its time variation occurs only at four times (t = O. tl. t2, t3). In between, it is independent of time, enabling us to use (2.64a) over the four lime intervals 0 10 tlo tl to t2, t2 to t3, t3 to t. Integralion of (2.116) across the discontinuities in 1{ shows thai ¢(t) is continuous. We use this fact to join solutions across the discominuities. Thus, using (2.120a) and (2.64a)
,
¢(tl) = ei,.lh 1\ Iz,p(O)
,
,Hltl
"'7r12
I I I--T--t I I I I
o
t,
(2.121.) (2.12Ib)
tz
etc.
Fig. 2.17. 11\ vcrsus time, giving the dcfinition of the tim.-s 0, '.,12, and 13 which specify thc beginning and end of the 1l/2 pulse (I 0, til and of the 1l pulse (t 12,13)
=
=
It is now convenient
10
define the quamities T(t. h o) and X(D) by (2.122.)
Then ¢ obeys the equation
a"ot
h-i
=
1e·¥'
where
X(O) = e
(2.116) (2.117)
Ito = Ho -w/-y
(2.118)
If , is positive, this will produce rotations about the effective field which are left handed. If, were itself negative, the rotations would be in the right-handed sense, as assumed in drawing Fig. 2.13. We do not set Ito '" 0 since we wish to represent the fact that owing to the inhomogeneity in }fo the typical spin is not perfectly at resonance. We introduce a distribution function P(1I 0 ) to express the fact that the number of spins with Ho between 11 0 and Ito + dho is
46
i81
"
(2.122b)
T(t. 110) generates the development of Ihe wave function during those times when HI = 0 for spins which are off resonance by " 0 . As can be seen by referring to (2.55), XeD) is the operator we need when we wish to rotate a componenl of spin through an angle D in the right-handed sense about the X -axis. Thus, if 10 becomes 10 , following such a rotation.
(2.123) For 0 = 1rn. X-I(1rI2)lyX(7r!2) = I z X- I (7r/2)lzX(7rf2.) = -ly
(2.124)
X~I(7r!2)I:r.X(7rf2.)= l oe
47
lf8=1r
X- I (1r)T- 1(t - T, ho)IyT(t - T, ho)X(1r) = X- I (1r)T- 1(t - T)X(7l') X- I (1r)lyX(1r)
X- 1(1r)Iy X(1r) = -Iy
,
X- 1 (1r)I.. X(1r) = -Iz
'--.-.2
... I
(2.125)
x ,X- I (1r)T(t - T, h o)X(7l') •
.
X- 1 (1I')I z X(1r) = I z .
(2.131)
3
Let us further consider that HI is so large that we can neglect the time intervals tl and t3 - t2 compared to T. Then, using expressions
We now have three tenns identified by the three numbered braces in (2.131). We deal first with term 2, using (2.125) to get
,,(II) = X(./2)"(0)
(2.126)
,,(12) = T(r, ho)X(.i2)"(O)
We write teml 3 as etc. we get ,pet) for times t which are after the second pulse as ,,(I) = T(I - r, ho)X(.)T(r, ho)X(.i2)"(O)
(2.132) (2.127)
for a single value of ho. Since the NMR signal arises from the transverse magnetization, we need to calculate the expectation value of I z and I y . We leave it as a homework problem to show that if H I lies along the x-axis. the expectation value of I z is zero as in the classical picture. Instead. we here calculate (Iy(t». For a single spin we get (2.128) where dTl stands for the "volume element" in spin space. Now, we must sum over all spins, using (2.119) to express the fact Ihat the various spins experience different ho's. Thus the expectation value of the total y-component of spin is (Iy,tota.](t» = N
J
P(ho)dho
j ,p*(ho, t)ly,p(ho, t) dT1
which is readily proved using the series expllllsion of the exponential and making repeated insertions of R- I R. Using (2.133), we then get X-I(1r)exp(iho(t - T)I.. )X(1r) =exp(iX- J (1r)I.. X(1r)ho(t - T» = exp(-i...,.ho(t - T)I.. )
J
,,'(O)X- I (.i2)T- I (r, ho)X-I(.)rl(1 - r, ho)IyT(1 - r, ho)
x X(.)T(r, ho)X(.i2)"(O)dr/
(2.130)
Utilizing the fact that X-I X = X X-I = I, we then transform a portion of the integrand:
(2.134b) We thereby get (Iy,tola.l(t» = - N
J
P(ho)dho
J
,p*(O)X- 1(1rI2)T- 1(T, ho)T(t - T, ho) (2.135)
Now consider what happens when t - T = T, i.e, t = 2T. This is the time of the echo according to our previous classical calculation. At this time so that
T-1(t-T,ho)T(T,ho)=1 (Iy,tolal(t = 2T» = -N
x 48
(2.134a)
x IyT-I(1 - r, ho)T(r, ho)X(.i2),,(O)dr/.
,p*(ho. t)Iy1/J(ho. t)dT[
=
(2.133)
(2.129)
where we have used a nalation ,p(ho·, t) which makes explicit the dependence of ,p on ho. We now utilize the properties of the exponential operators [see (2.50a)] to transform the integral over T1 10 get
J
This term has I z in the exponent. To deal with it, we call on a theorem (given as Problem 2.4) that if R is an operator, R- l its inverse, and G some other operator
(2.136)
J
p(ho)dho
J
,,·(O)X- I (.i2)Iy X(./2),,(O)dr/
(2.137) 49
The ho has now disappeared from the integral over dr/, so that we can now integrate over dho, giving (2.138) The integrand is just what we would have if we calculated (1",lolal) immediately after the first 1r(1. pulse. Using (2.124), in fact, we have (2.139) where 0- refers to the time just before t = O. If we assume the system is in thennal equilibrium before the r(1. pulse, this resull states that the echo arises from the thennal equilibrium Mo. Note that the minus sign shows that the echo forms on the negative y-axis. This same result would follow from the classical argumenl for a nucleus with a positive.., for which the rouuions about HI are in the lefl·handed sense. Going back to (2.134) and (2.135), we see that the effect of the 1r pulse may be thought of as having changed the Hamihonian after t = T from its old value of "Hold to a new value "H new . In this viewpoinl, Ihe Hamiltonian from tl to t2 is "Hold = -..,llholn and then at t = r changes to its negative, "Hn(~w = +-,1I11 01z for later times. We replace the 1r pulse plus subsequent evolution under the real Hamiltonian with a developmenl under a new or effective Hamiltonian for t > T in which there is no 11' pulse. The effect of the pulses is equivalent to a situation in which there aTe no pulses. bllt the Hamiltonian changes in time. This concept proves very useful in magnetic resonance. II is the basis of many important pulse sequences. In the case of the echo, the combination of 'H new and "Hold is such that when t = 2T, 'H new has completely undone the effect of 'Hold. Over the time interval 2T we have eliminated the effect of magnetic field inhomogeneities from the Hamiltonian. The echo can be used to eliminate, over a particular time interval, the effect of any interaction which is equivalent to a spread in magnelic field. For e)(ample, spin-spin couplings 10 different nuclear species (e.g. coupling of protons to C 13 , when Ct3 is under observation) or the effects of chemical shifts discussed in Chap. 4 all can be eliminated by use of echoes. By using Ihe explicit expression of (2. I34a), it is easy to show that apart from a sign change the development of (ly,lotat(t» for t > 2T is identical to its development immediately after the first 11'!2 pulse. We give this as a homework problem. This result has a useful e)(perimental consequence. The signal immediately after the 7r!2 pulse gives one the Fourier transform of the line shape function P(ho), as can be seen from the results of Problem 2.3. But the signal immediately follows a strong pulse. In practice this pulse may block the signal amplifiers of the NMR apparatus, making it impossible to observe the signal until the amplifiers recover. On the other hand, the echo occurs later than a pulse by a time T, giving the amplifiers time to recover. Thus, the echo is easier to see. It is therefore fortunate Ihat the echo reproduces the earlier signal. 50
2.11 Relationship Between Transient and Steady-State Response of a System and of the Real and Imaginary Parts of the Susceptibility Suppose, to avoid saturation, we deal with sufficiently small time-dependent magnetic fields. The magnetic system may then be considered linear. That is, the magnetization produced by the sum of two weak fields when applied together is equal to the sum of the magnetization produced by each one alone. (ytIe shall not include the slatic field Ho as one of the fields, but may find it convenient to consider small changes in the static field.) In a similar manner, an ordinary electric circuit is linear, since the current produced by two voltage sources si· multaneously present is the sum of the currents each source would produce if the other voltage were zero. Let us think of the magnetization Ll."f(t) produced at a time t and due to a magnetic field H«() of duration Llt' at an earlier time (see Fig.2.18). As a result of the linearity condition we know that LlM(t)oc H(t'). It is also oc Llt' as long as Llt' <: t - t' , since two pulses slightly separate in time must produce the same effect as if Ihey were applied simultaneously.
H(t')
•
Fig.1.18. Pulse or magnetic field
Therefore we may express the proponionality by writing LlM(t) = met - t')H(t')Llt!
(2.140)
where met -t' ) is a "constant" for a given t and t ' , which, however, must depend on how long (t - t' ) after the pulse of field we wish to know the magnetization. The total magnetization at time t is obtained by integrating (2.140) over the history of the magnetic field H(t'):
,
M(t) =
J
met - t')H(t')dt'
(2.141)
-= Note that met - t') = 0 if t' > t, since the effect cannot precede Ihe cause. To understand just what met - t') is, lei us assume H(t ' ) is a 6-function at t = O. Then the magnetization at t > 0 (which we shall denote by M6) is 51
,
M6(t):::
J
met - t')E(t')dt'
=- met)
(2.142)
.
-~
That is, met) is the response to as-function al t ::: O. Knowledge of met) enables us to detennine from (2.141) the magnetization resulting from a magnetic field of arbitrary time variation. Ordinarily one detennines met) by going 10 the basic lime-dependenl differential equation which describes the behavior of the system. We can illustrate the approach by an example. Suppose we consider a system described by (2.81)
with Mo = XoHo· We ask what the effect would be if Ho were taken to be time dependent in magnitude though tilted in direction. We replace Ho by H(t). This silUation describes the famous experiments of CJ. Gorttr and his colleagues at [.e;den [2.18).
Those readers familiar with Green's functions will recognize that met) is a Green's function, and that what we have just done is one example of how one finds a Green's function . .If a unit step were applied at t 0 (Fig. 2.19), we should have magnetiuttion, which we shall denote as M slep :
=
Mslep(t) =
(2.143)
dm m XO - + - = -'(I) dt T1 T1
(2.144)
We know from causality that met) = 0 for t when the right·hand side is zero, the solution is
< O.
We know that for t
met) = Ae- I/ T1
> 0,
J
dm +
1=0-
J~
dt =
1=0-
~~
d met) = dt (Mstep)
(2.146)
t=O-
(2.150)
Equation (2.150) therefore shows us that knowledge of Mstep(t) enables us to compute met). For example, suppose we discuss the magnetization of a sample following application of a unit magnetic field in the z-direction for a system obeying the Bloch equations. We know from the Bloch equations that
= Mlltep
(2.151)
Therefore, using (2.150),
T,
E(t)dt
Fig. 2.19. Step runctKm
By taking the derivative of (2.149), we find
1=0+
J
(2.149)
0
,
met) = XOe-t/TI 1=0+
j m(r)dr
o
MzCt) = xo{l- e- t / T1 ) (2.145)
To find A, we integrate (2.144) across t = 0 from t = 0- to t = 0+:
t=O+
t')dt' =
_I_H(O_ _
To find the equation for m{t), we recognize that met) obeys this equation when H(l) • '(I). Thus.
j met o
We then have
dM% + M r = XO H(t) dt T1 Tl
,
,
(2.152)
Note that in any real system, Ihe magnetiutlion produced by a step is bounded, so that ~
J
which gives
m(r)dr
m(O+) _ m(O-).
<0
T,
T,
o
(2.147)
converges. Suppose we apply an alternating magnetic field. We shall write it as complex for simplicity:
Since m(O-) = 0, and m(O+) = A [from (2.145)], we get A = XO/TI, or met) = XOe-t/Tt
(2.153)
(2.148)
H~(t) = H Xoeil
(2.154)
Then 52
53
,
M~(t)
J
:::
1
m(t - t')HXOeiwt' dt'
mer) = 211"
,
-00
J met -
= HXOeiwt
-00
-00
J
m(r)e-I"'T dT
(2.155)
o Comparison with (2.96) shows that
J 00
X=
m(r)e-i"'T dr
o
x'= jm(T) cos wTdr
(2.156)
o
Jri
J
00
XII =
meT) sin wrdr
o (see
That"h, m(r) and X(w) are Fouricr transfonns of each other. Knowledge of one completely delennines Ihe other. One may attempt to predict the properties of resonance lines either by analyzing the response to an alternating signal or by analyzing the transient response. Kubo and Tomita [2.19J, for example, base their general theory of magnetic resonance on the transient response, calculating the response of the system to a step. Examination of (2.156) enables us to say something about X' and i' at both zero and infinite frequencies. Clearly, X/l vanishes at w = 0, since sinO vanishes, but x' does not vanish at w = O. Moreover, if mer) is a finite, reasonably continuous function whose total intcgral oo m(r)dr is bounded, both X' and X" will go to zero as w ----+ 00, since the oscillations of the sin wr or cos wr will "average" the integrand to zero. Actually we may pennit m(r) to be infinite at r = O. We can see this by thinking of m(r)dr, the response to a step. We certainly do not expect the response to a step to be discontinuous at any time other than Ihat when the step is discontinuous (t = 0). Therefore m(r) can have at most an integrable infinity at t '" 0, since the response must be bounded. We shall represent this by a a-function. Thus, if
Jo
00
,,
(2.158)
x(w)ei,,",(dw
t')eiw(tl-t)dt'
00
= Hxoeilo!t
J
+00
footnote~).
It is simple to show, using the integral representation of the
c-funclion, 1 6(x) = 21r
J
mer) = m,(r)
+00
that
eixtdt
(2.157)
+ Clo(r)
(2.159)
where m,(r) has no a-function, we gel
-00
Jm,
00
X/(w) = S Strictly speaking, we should turn on the alternating field adiabatically and consider the limit of slower and slower turn-on. Thus we can lake
fl~(t) =IJxoeiw,c'"
6>0
As t ..... - 00, this function goes to zero. We compute the limit as 8 ..... O. Thus
M~(t)
•
=
J
(2.160)
c,
The integral vanishes as w ----+ 00, leaving us = x'(oo). It is therefore convenient 10 subtract the o-funclion pan from m(r), which amounts 10 saying that
J
00
X(w) - X'(oo) =
met - t')Hxoc i .." " e"' dt'
-00
= lI xo e'W'c"
J J
met - t')eiw("-')c'{"-l)dt'
00
= Ilxoe(iw+.),
m(r)e-(·+iwjT dr
m(r)e-i,,",T dr
(2.161)
o
•
_00
and
•
.-.
x("') = Jim jm(T)e-(.+'W)Tdr The advantage of this definition is that it has meaning for the case of a ~lossless resonator" (magnetic analogue of an undamped harmonic oscillator), in which a sudden application of a field would excite a transient that would never die out.
54
(1') cos (wr)dr + c\
o
where now m(r) has no a-function part. [Of course no physical system could havc a magnetization that follows the excitation at infinite frequency. However, if one were rather making a theorem about penneabilitY!1-, p(oo) is not zero. We keep x'(oo) to emphasize the manner in which such a case would be treated.] We wish now to prove a theorem relating X' and X", the so-called KramersKronig theorem. To do so, we wish to consider X to be a function of a complex variable z = x + iy. The real part of z will be the frequency w, but we use the symbol x for w 10 make the fonnulas more familiar. Therefore
55
=
x(z) - i(oo) =
=
J o = J
m(r)e- iZT d,
The presence of the tenn exp (yr) tells us that Ix(z) - i(oo)I--+ 0
m(T)eYTe-i:l;T dT
(2.162)
o Since an integral is closely related to a sum, we see that X(z) is essentially a sum of exponentials of z. Since each exponential is an analytic function of z. so is the integral, providing nothing too bizarre results from integration. To prove that X(z} - x'(oo) is an analytic function of z, one may apply the Cauchy derivative test, which says that if x(z) - X/Coo)
== U + iv
(2.163)
and
avail ax = - ay
(2.164)
From (2.162) we have
J= o = J
m(T) cos (xT)e
00
-
·We already know that
Ix(z)-i(oo)I-O as
X--I±OO
Therefore X(z) - i(=) is a function that is analytic for Y:5 0 and goes to zero as Iz I -+ 00 in the lower half of the complex plane. Let us consider a camour integral along the path of Fig. 2.20 of the function
xCi) - x'(oo) z' w
X'(z') ~ X'(oo)
j "--''--'-;---'''-'.:.'-'dz
I
ZI_W
u ""
Y --I
By Cauchy's integral theorem this integral vanishes, since X(z) has no poles inside the contour.
where u and v are real, u and v must satisfy the equations
au au ax = ay
as
=0
(2.167)
C
yT
(2.165)
dT
Since IxIV) - x/(oo)1 goes to zero on the large circle of radius e, that part of the integral gives zero contribution. There remains the contribution on the real axis plus that on the circle Zl - w = Rexp (i¢). Thus
m(T) sin (xT)eyTdT
v = -
o giving
=
+
au = -jm(T)T sin (XT)eYTdT= av
ax
au
-
ay
o
ay
ax
which satisfy the Cauchy relations, provided it is pennissible to take derivatives under the integral sign. There are a variety of circumstances under which one can do this, and we refer the reader to the discussion in Hobson's book [2.20]. For our purposes, the key requirement is that the integrals in both (2.165) and (2.166) must not diverge. This prevents us in general from considering values of y that are too positive. For any reasonable mer) such as that of (2.152), the integrals will be convergent for y:S 0, so that X(z) - X'(oo) will be analytic on the real axis and in the lower half of the complex z-plane. Whenever we use functions mer) that are flOl well behaved, we shall also imply that they are to be taken as the limit of a well-behaved function. (Thus an absorption line that has zero width is physically impossible, but may be thought of as the limit of a very narrow line.)
56
Re1q,
= O= P
j
-=
+=
j
o:.J+R
11"
+00
j m(T)r cos (XT)eyTdr = -a,o
'J
,
X(w) -.X (00) Rieiq,d¢+
(2.166)
= =
j
2. [
,
I
X(w') -l(oo)
wI-w
, )J w + 1T1.[ X(W ') -X(oo
X(w) - X (oo)d'
w
I
w
(2.168)
where the symbol P stands for taking the principal pan of the integral (that is, taking the limit of the sum of the integrals ~~R and J:;+~ as R --+ a simultaneously in the two imegrals).
Fig.2.20. Contour integral 57
x"
Solving for the real and imaginary parts, we find
J
1 +00 1/( ') x(w) - x'(oo) = -p ~c/w' 11" w'-w
-n
(2.169)
-00
+oox" (w ) - X,(00) dw' X"(w) = _2.. p
J
(a)
n
w
I I
n!(
~-n
-00
These are [he famous Kramers-Krollig equations. Similar equations can be worked out for analogous quantities such as the dielectric constant or the electrical susceptibility. The significance of these equations is that there are restrictions placed, for example, on the dispersion by the absorption. One cannot dream up arbitrary X'(w) and X"(w). To phrase alternately, we may say that knowledge of X" for all frequencies enables one to compute the X' at any frequency. Note in particular that for a narrow resonance line, assuming x'(oo) = 0, the static susceptibility Xo is given by I
XO = X(O) = -p •
J _,_w_
+00
(b)
I I I
I I
w'-w
1f
.
X
I I I
w
fig. 2.11. (a) Absorption spectrum. (b) Corl"C$ponding dispel'!lion spedrum
,
,
x (W)-X(OO)
c( 1
=11"
I) c(- -I + -I) -
--f}-w -f}-w
=11"
fl-w
fl+w
(2.172) where we have used the fact that 6(%) = 6(-%).7 Of course, near resonance (w ";l! fl), only the firsl tenn is large. The function is shown in Fig. 2.21.
..fl( ')
2.12 Atomic Theory of Absorption and Dispersion
w'
-00
-
2 I +00
- •-wo -
J"( x ')"
(2.170)
wuw
o
The integral of X"(w l ) is essentially the area under the absorption curve. We see that i[ may be computed if the static susceptibility is known. G As an example, suppose
m- 6(-w -
x"(w) = c[6(w -
fl)]
(2.171)
The first tenn corresponds to absorption at frequency n. The second tenn simply makes x" an odd function of w. For this funclion, whal is X'(w)? x'(w) - X'(oo)
=
2.. p
J c{6(w' - m- 6(-w' - n>jdw'
+00
wI-w
7l"
-00
G or course, if we arc talking about a m... gnetic re>;onance exp·eriment with the static field in ~he z-direc~ion and the alternating field in the z-direetion, we are discussing Xu. Then x'(O) or (2.170) is x~...(O), whereas xo is usually thought or as relating the total magnetization Alo to the field I/o, which produces it, and is thllS X~r{O). lIowever, a Slnll.lIstatie field IIr in the z-directioll simpl)' rotates M o , giving
I/r M oII,
,
=
x.. (0
ThllS X~...(O) = x~AO)
=Xo.
Mr =
58
)
lIr
We shall now tum to obtaining expressions for the absorption and dispersion in tenns of atomic properties such as the wave functions, matrix elements, and energy levels of the system under study. We shall compute X" directly and obtain X' from the Kramers-Kronig equations. We make the connection between the macroscopic and the microscopic properties by computing the average power P absorbed from an alternating magnetic field HrlJ cos wt From (2.102) we have
P -w"H'V -"2X zO
(2.173)
in a volume V. It will be convenient henceforth to refer everything to a unit volume. (We shall have to remember this fact when we compute the atomic expressions in particular cases). On the other hand, Ihe alternating field couples 10 the magnetic moment Itzk of the klh spin. Therefore. in our Hamiltonian we shall have a time-dependent perturbation 1{t)er~ of 7 To show that 6(z) = 6(-z) we consider the integrals II = I~oo f(z)6(z)dz and =I~::: f(z)6(-z)dz. We have irnmedia~ely that II = j(O).1b eval~ate 12, we chAnge variable to z' = -z. Then we get
I~
+12 = Thus 12
J
--
f( -z')6(z')dz'
=/(0)
= II, and ,s(z) =6(-:r:). 59
'}-{pert = - L!1-3:kH3:0 cos wt k = -!1-1;H1;O cos wt
(2.174)
We can compute the absorption rate P ab, due to transitions between states a and b in tenns of Wah, the probability per second that a transition would be induced from a to b jf the system were entirely in state a initially:
where !1-3: is the x-component of the total magnetic moment ~§L!1-1;k
(2.179) a.175)
k
In the absence of the perturbation, the Hamiltonian will typicnlly consist of the interactions of the spins with the external static field and of the coupling '}-{'k . . J between SpinS J and k. Thus
'H" - L"'kHO+ L'Hjk k
(2.176)
The tenns P(Eb) and P(Ea ) come in because the states la) and Ib) are only fractionally occupied. The calculation of the transition probability Wah is well known from elementary quantum mechanics. Suppose we have a time-dependent perturbation Hpert given by HI>ert = Fe -iw! + Ge iw !
(2.180)
j,k
We shall denote the eigenvalues of energy of this many-spin Hamiltonian as E a , Eb, and so on, with corresponding many-spin wave functions as la) and Ib). See Fig.2.22. Because of the large number of degrees of freedom there will be a quasi-continuum of energy levels.
where F and G are two operators. In order that Hpert will be Hennitian, F and
G must be related so that for all states I(t) or Ib), (aIFlb) " (bIGla)'
(2.181)
Under the action of such a perturbation we can write that Wah is time independent and is given by the fonnula (2.182)
Fig.2.22. Eigenvalues of energy
The states la) and Ib) are eigenstates of the Hamiltonilln. The most general wave function would be a linear combination of such eigenstates: (2.177)
where the ca's are complex constants. The square of the absolute value of Ca gives the probability p(a) of finding the system in the eigenstate a: p(a)"
1,.1 2
If the system is in thennal equilibrium, all stales will be occupied to some extent, the probability of occupation p(a) being given by the Boltzmann factor
e- EA / kT
P(E.) "
Le
E,/kT
h l(a!Flb)1 < T
(2.178)
E.
where the sum E c goes over the entire eigenvalue spectrum. The denominator is just the classical partition function Z, inserted to guarantee that the total probability of finding the system in any of the eigenstates is equal to unity; that is,
60
provided certain conditions are satisfied: We do not ask for details that appear on a time scale shorter than a certain characteristic time T. It must be possible to find such a time, which will satisfy the conditions that 1) the populations change only a small amount in T and 2) the possible states between which absorption can occur must be spread in energy continuously over a range LJ.E such thaI LJ.E» hiT. These conditions are violated if the perturbation matrix element l(aIFlb)1 exceeds the line width, as it does when a very strong alternating field is applied. We can see this point as follows: The quantity LJ.E may be taken as the line width. We have, then, that LJ.E< l(alFlb)l. But under these circumstances one can show that the populations change significantly in a time of order IiII(alFlb)j. Thus to satisfy the condition I that the populations change only a small amount during T, T must be chosen less than hll(aIFlb)l. This gives us
But, by hypothesis,
dE < l(alFlb)1
Therefore
LJ.E < !!. T
which violates condition 2 above. Thus it is not possible to satisfy both conditions, and the transition probability is not independent of time. 61
,
This example shows why we did not gel a simple time-dependent rate process in Sect. 2.6, since for that problem, the energy levels in the absence of HI are peneClly sh"'P (LIE = 0), l(aIFlb)1 > LIE. In our formula for Wah we use the O-funclion. This implies thai we shall evenlUally sum over a quasi-continuum of energy states. In writing the rransition probability, it is preferable to use the o-function (onn rather than the integrated form involving density of slales in order to keep track of quantum numbers of individual stales. By summing over all states with E a > Eb, we find
21f H 2o 2 -'-r,w !P(E,) - p(E,)JI(al",lb)! ,(E, - E, - I>w) 11 4 Ea>Eb
The quanta hw correspond crudely to the energy required to invert a spin in the static field. This energy is usually much smaller than kT. For nuclear moments in strong laboratory fields (...... 10,1 Gauss), T must be as low as 10- 3 K so that fiw will be as large as kT. This fact accounts for the difficulty in producing polarized nuclei. For electrons, kT ...... fiw at about 1 K in a field of 104 Gauss. Therefore we may often approximate (2.188) We may call this the "high-temperature approximation". By using (2.178) and (2.188), we have
L:
P = -
=~i'H;o
= e-Ea/kT[e(Ea-Eb)/kT _ I]
P(E,) - p(E,)
(2.183)
Z =
e- Z (Ea kT- Eb) Ea kT /
(2.189)
Therefore
I
X"(w)
=, L:
[P(Eb) - p(E,)II(a!",lb)I',(E, - E, - hw)
. (2.184)
Ea > Eb
Eb.
As long as Ea > only positive w will give absorption because of the 0function in (2.184). Removal of the restriction E a > Eb extends the meaning of xll(w) fonnally to negative w. Note that since lJ(Eb) - P{Ea ) changes sign when a and b are interchanged, i'(w) is an odd function of w, as described in the preceding section: X"(w)
=, L:
IJKEb) - p(E,)II(al,',lb)I',(E, - E, - I>w)
.
(2.185)
Ea,Eb
Assuming i(oo) = 0 for our system, we can easily compute i(w), since x'(w)=.!.P
+=
Jw
1T
lie
I
XI w)dw ' -w
(2.186)
-00
J ,(E, -,Eo -
+00
=, L:
[P(E,) - P(E,lll(al",lb)I'.1. p
Ea,Eb
1T
W
-00
-
,
hw) dw'
W
or, evaluating the integral, x'(w) =
L:
[P(E,) -
P(E,)JI(a!",lb)1
Ea,Eb
,
1
E _ E _, a
b
(2.187)
lW
By lIsing the fact that a and b are dummy indices, one may also rewrite (2.187) to give x'(w) =
L: p(E,JI(al,',lb)I' [(E, -
E, - nw)-l
Ea,Eb
+ (Ea 62
- Eb
+ rlW)~I]
Substitution of (2.189) into (2.185), together with recognition that E a - E b = hw, owing to the o-functions, gives xl/(w) =
~;; L:
e-Ea/kTI(alfl':rlb)120(Ea - E b - hw)
(2.190)
Ea,Eb
Another expression for XI/(w) is frequently encountered. It is the basis, for example, of Anderson's theory of motional narrowing [2.21]. We discuss it in Appendix B because a proper discussion requires reference to some of the material in Chapters 3 and 5. It is important to comment on the role of the factors exp(-Ea/kT). If one is dealing with water, for example, the proton absorption lines are found to be quite different at different temperatures. Ice, if cold enough, possesses a resonance several kilocycles broad, whereas the width of the proton resonance in liquid water is only about 1 cycle. Clearly the only difference is associated with the relative mobility of the H 20 molecule in the liquid as opposed to the solid. The position coordinates of the protons therefore play an important role in detennining the resonance. Fonnally we should express this fact by including the kinetic and potential energies of the atoms as well as the spin energies in the Hamiltonian. Then the energies E a and Eb contain contriblllions from both spin and positional coordinates. Some states la) correspond to a solid, some to a liquid. The factor exp(-Ea/kT) picks out the type of "lattice" wave functions or states that are representative of the temperature, that is, whether the water molecules are in liquid, solid, or gaseous phase. Commonly the exponential factor is omined from the expression for XII, but the states la) and Ib) are chosen to be representative of the known state. The classic papers of Gutowsky and Pake (2.22], on the effect of hindered molecular motion on the width of resonance, use such a procedure.
(2. 187a)
63
Evaluation of x!' by using (2.190) would require knowledge of the wave functions and energy levels of the system. As we shall see, we rarely have that infonnation, but we shalI be able to use (2.190) to compute the so-called moments of the absorption line. We see that the only frequencies at which strong absorption will occur must correspond to transitions among states between which the magnetic moment has large matrix. elements.
3. Magnetic Dipolar Broadening of Rigid Lattices
3.1 Introduction A number of physical phenomena may contribute to the width of a resonance line. The most prosaic is the lack of homogeneity of the applied static magnetic field. By dint of hard work and clever techniques, this source can be reduced to a few milligauss out of 104 Gauss, although more typically magnet homogeneities are a few tenths of a Gauss. The homogeneity depends on sample size. Typical samples have a volume between 0.1 cc to several cubic centimeters. Of course fields of ultrahigh homogeneity place severe requirements on the frequency stability of the oscillator used to generate the alternating fields. Although these matters are of great technical imponance, we shall not discuss them here. If a nucleus possesses a nonvanishing eleclric quadrupole moment, the degeneracy of the resonance frequencies between different m-values may be lifted, giving rise to either resolved or unresolved splittings. The latter effectively broaden the resonance. The fact thaI T, processes produce an equilibrium population by balancing rates of transitions puts a limit on the Iifctime of the Zeeman stales, which effectively broadens the resonance lines by an energy of the order of lilT,. In this chapter, however, we shall ignore all these effects and concentrate on the contribution of the magnetic dipole coupling between the various nuclei to the width of the Zeeman transition. This approximation is often ex.cellent, particularly when the nuclei have spin! (thus a vanishing quadrupole moment) and a rather long spin-lattice relaxation time. A rough estimate of the effect of the dipolar coupling is easily made. If typical neighboring nuclei are a distance r apan and have magnetic moment lA, they produce a magnetic field Bloc of the order I' Bloc ::; r 3
(3.1)
By using,. ::; 2 A. and J-l ::; 10- 23 erg/Gauss (10- 3 of a Bohr magneton), we find Hloc s::o I Gauss. Since this field may either aid or oppose the static ficld Ho, a spread in the resonance condition results, with significant absorption occurring over a range of B 0 ....., I Gauss. The resonance width on this argument is independent of H 0, but for typical laboratory fields of 104 Gauss, we see there is indeed a sharp resonant line. Since the width is substantially greater than the magnet inhomogeneity, it is possible to study the shape in detail without instrumental limitations. 64
65
Flg.3.t. Relationship between rectangular coordina~es :1:, II, , (describing the position of nucleus 2 rdative to nucleus I) and the polar coordinates r, 8, '"
z
3.2 Basic Interaction The classical interaction energy E between two magnetic moments III and 112
;s (3.2)
k--li-r7-- -"
where r is the radius vector from III to JJ.2. (The expression is unchanged if r is Iaken as the vector from 112 to PI.) For the quanlum mechanical Hamiltonian we simply take (3.2). treating 1-'1 and P2 as operators as usual:
(3.3)
P2""'Y2hI2.
PI ""/lhII
where we have assumed that both the gyromagnetic ratios and spins may be different. The general dipolar contribution to the Hamillonian for N spins then becomes
Hd ""
~
EE
[Ili
2 j=1 1e=1
~PIe
_ 3(pj
r jle
.rjk~(Il/c .rjlIJ]
(3.4)
rjk
where the! is needed. since the sums over j and k would count each pair twice, and where. of course, we exclude tenr.s with j = k. By writing PI aod JJ.2 in component fonn and omitting the subscriplS from r, we see from (3.2) that the dipolar Hamihonian will contain terms such as 2
I
2
zy
/112k [lzh="5
/112h hzI2z-, r
r
r'
/12
(A + B + C + D + E + F)
r:-
(3.6)
y
z
Hz "" -71hRoIII - 72hHo12z corresponds to an interaction with a field of 104 Gauss. It is therefore appropriate to solve the Zeeman problem first and then treat the dipolar term as a small penurbalion. (Actually. for two spins of~, an exact solulion is possible.) To see the significance of the various tenns A, B, C. and so on, we shall consider a simple example of two identical moments. both of spin The Zeeman energy and wave functions can be given in tenns of the individual quantum numben ml and m2. which are the eigenvalues of II' and h,· Then the Zeeman energy is
!.
(3.8) We shall diagram the appropriate matrix elemenlS and energy levels in Fig.3.2. It is convenient to denote a state in which ml = m2 = -~ by the nOlation 1+ -). The twO states 1+ -) and 1- +) are degenerate. and both have B-l = O. The states 1+ +) and 1- -) have, respectively. -hwo and +hwo, where wo = ,Ro as usual. We first ask what pairs of states are connected by the various tenns in the dipolar expression. The tenn At which is proponional to II.zh.z' is clearly completely diagonal: It connects Iml m2) with (mtm21. On the other hand, B, which is proportional to It 12" + II only connects Iml m2) to states (ml + I, m2 - II or (ml - I, m2 + II. A customary parlance is 10 say that B simultaneously flips one spin up and the other down. B therefore can join only the states 1+ -) and 1- +). The states joined by A and B are shown diagrammatically in Fig. 3.3.
+!,
It,
where 2
A"" [1,[2,(1 - 3 cos 8)
B = -t(I{ 12" + I) [i)(I
C=
1/
(3.5)
If we express 1)z and [151 in tenns of the raising and lowering operators and I), respectively, and express the rectangular coordinates z, y, z in tenns of spherical coordinates r, 8, tP (Fig. 3.1), we may write the Hamiltonian in a fonn that is particularly convenient for computing matrix elements: '}-{d = /11'2
I'
- 3 cos 2 8)
-~(It h, + 1),Ii) sin B cos Be-i,p
(3.7)
D = -~(Il h, + IlrI2") sin 8 cos 8e i,p E= -~I{Ii sin 2 8e- 2i ,p . 2 ee 2i ,p F -- -;{'1-1I 2 Sin As we have remarked, (/11'2h2)1r3 corresponds to the interaclion of a nuclear moment with a field of about I Gauss, whereas the Zeeman Hamillonian
66
-----+ +--------++ Fig. 3.2. Energy levels or two
iden~iclIl
o
spins
67
,--A ...
(
Fig. 3.3. States joined by matrix elements A .. nd B. The dashed lines &0 between states thal are joined
)
A
...-........ +- ( ,
-----_ .... B
A ....-.... >
( )
.....
where
is the wave function corrected for the effect of the penurbation 'Hpert
the unpenurbed states of u~, and u~. Qy means of (3.9) we can see that the state 1+ +) will have a small admixture of 1+ -), 1- +), and 1- -). The amount of admixture will depend on (n'I'Hpertln) and En - En" The former will be -y 2h 2Jr3 multiplied by a spin matrix element. Since the spin matrix element is always of order unity, and since Hl oc = -yTl/r 3, we can say (n/l'Hperdn) ':! -yIIHloc . On the other hand, En - En' = Ilwo = 'YfIHO, so that
-+
,..-_ ....A
'+ +
(
Un
and where, of course, the matrix elements (n'j'Hpertln) are computed between
Note that B has no diagonal matrix elements for [he m 1m2 representation, but it has off-diagonal elements between two stales which are degenerate. The fael that off-diagonal elements join the degenerate states 1+ -) and J _ +)
(n'I1ip<,d nl En - En'
I
tells us, of course, that they are not the proper zero-order states. B therefore
I
(3.IOl
corresponding to a very small admixture. Of course the admixture produces a second-order energy shift. As a second, and for us more important, effect, the admixture enables the alternating field to induce transitions that would otherwise be forbidden. Thus the transition from 1+ +) to 1- -), which would be forbidden if these were exactly the states, can now take place by means of the small admixture of the states 1+ -) and 1- +). (See Fig. 3.6.)
plays an important role in determining the proper zero-order functions. When Ihe ~roper zero-order functions are delcrmined, B turns QuI to have diagonal matnx elements. We shall return to Ihis point later. Since terms C and D each flip one spin only, they join states shown in Fig. 3.4. all of which differ by liwo in energy. Finally E and F flip both spins up or both spins down, connecting states that differ by 2Tlwo (Fig. 3.5).
++--.~,
+-
+----
-+
-"-----+
Fig. 3.6. The strong transition is indicated by the double arrow. The transition by ~he tight arrOW has non· vanishing matrix clements due to the dipole I\dmix~urcs
Fig.3.4. States joined by the terms C aud D
The matrix element is smaller than the normal one, for example, between
1+ +) and 1+ -), in the ratio H1oc1Ho. 1llerefore the intensity of the absorption.
i
+----
lE,F
+
I
I I I
++
Fig. 3.5. States joined by the dipolar terms E and F
.The terms C, D, E. and F therefore are off-diagonal. They produce slight admixtures of our zero-order states into the exact stales. The amount of the admixture can be computed by second-order penurbation Iheory. using the wellknown expression for the correction of the zero-order functions U?I of zero-order energy En: U
= u O + '" (n'l1-lperdn) 0 nnL..EEun' n'
68
n
n'
(3.9)
,
which goes as the square of the matrix element, is weaker in the ratio (HlocIHo)2. The transition occurs, of course, at w = 2wo. A funher consequence of the admixture of states is that a transition near w = 0 can be induced. [Actually this transition is forbidden for a pair of spins, each of spin!, because the eigenstates of }.II = mt + 1n2 = 0 are of different symmetry under exchange of the panicle labels ((he singlet and triplet states), whereas the penurbation is symmetric. If more than two spins are involved, the transition is permilled.] The net effect of the terms C, D, E, and F is therefore to give the absorption near 0 and 2w0, shown in Fig. 3.7. The extra peaks at 0 and 2wo are very weak and may be disregarded for our purposes. Since they are the principal effects of the terms C, D, E, and F, it will be an excellent approximation to drop C, D, E, and F from the Hamiltonian. For some of our later calculations we shall see that failure 10 drop these terms can lead us into erroneous results. The remaining dipolar term A + B may be combined to give what we shall call
Jt3:
69
Ab."pti,"
l_-"--,-,,fl_~L"":"'..~ 140
2"'0
3.3 Method of Moments' Before outlining the method of moments, we must return to our original expression for xl/(w):
'"
Fig.3.7. Absorption VCMlUII frequency, including dipolar couplings. The three absoq)~ion reg;ona h....e width '" -rlllo~, but the intensity of the peaks at 0 and 2lJg is...., (1l Ioc lllo )2 smaller than that al lJg
(3.11) and the total simplified Hamiltonian becomes I
1i = L(-,1l.HOlz /r.) + 4,2h. 2 k
2
9'1:) E (1-3005 3 ) (3IjzI/cz i,k
rjA:
Ij' 1k) (3.12)
Now the tenns 1{z and ~ commute. (This can be seen by considering a pair of spins I and 2. Clearly I:: = II: + 12z commutes with 31):12::' How about the tenns I) . h? I: commutes with (II + 12)2, since I) + 12 = 1 is the operator of the total angular momentum (any component of angular momentum commutes with the square of the total angular momentum). By writing (l '" (I) + 12)2, we
have P=Ir+l~+2II·I2
(3.13)
We see thai Ih + h: commutes with the left side and the first two tenns on the right. Therefore it must commute with II . h.] If two operators commute, we may choose an eigenfunction to be simultaneously an eigenfunction of both. Let us use a to denote the eigenvalues of~. Then we have
x"(w):
JiZUM::
(--yllHoM)UM
so thaI
(Jiz + Ji3)IMa):: (-'YhHoM + Eo)IMa)
(3.15)
These quantum numbers will prove useful later. Unfortunately all we can say about the quantum numbers a is that they exist, although we do not know them or the corresponding eigenfunctions. If Ji~ consisted of only the term IltI2:' we could solve the resonance shape exactly. We could do the same if all we had was It . I2. But the presence of both together, since they do not commute, spoils the two solutions. In fact, to proceed further, we are forced to go to the so-called method of moments, a clever technique due to Van Vleck, which enables one to compute properties of the resonance line without solving explicitly for the eigenslates and eigenvalues of energy.
70
.,'
E, -1>w)
(3.16)
Since we shall treat the lauice variables as parameters, the only variables coming into the problem are due to spin; thai is, the quantum numbers a and b refer to spins. We shall therefore assume Ell
!'
-IN
j(w): (3.14)
:~"; Le- E./kr l(al".lbJI'5(E. -
L l(al,'.lb)I'5(E. -
.,'
(3.17)
E, - low)
In fact, experimental determinations of Xl/(w) enable us to compute few) from (3.17), and conversely a lheoretical dctennination of few) gives us xl/(w). We focus therefore on few). First we note that since X/I(W) was an odd function of w, few) is an even function. [This fact is also evident by explicit examination of f(w).] We now define the nih moments of f(w) by the equation
=
J (wn) :: __ Jj(w)dw w n f(w)dw
~o"'=;-
and
(3.18)
o
See rererences under "SccorKI
MOlllcnt~
in
~he
Bibliography 71
~
So, we can rewrite (3.23) as
j(w - (w»n j(w)dw (&.,n) =
~O'---O~,,
_
(3.19)
o = 2 is called the "second moment". Clearly (L1w 2 )
The expression (3.19) for n is of the order of the square of the line width, so that
(3.20)
The two moments of (3.18) and (3.19) are closely related. as may readily be seen = w 2 - 2w(w) + (w 2 ), one easily
for n:: 2 as follows: Expanding (w - (w»2 shows from (3.18) and (3.19) that
.
(3.21)
k
j(w)dw. which IS, of course, closely related 10 the area under the absorption curve. Then we shall compute (w) and (L1w2). Since f(w) is an even function, To illustrate the general methods, we shall first campUle
~
j j(w)dw =
o
i
+~
j
j(w)dw
=2
j D al".lb)(bl".)a)6(E. -
E. - hw)dw
(3.22)
-00 G,b
The integrand picks up a contribution from the IS-function integral every lime hw = E a - E b- But for any pair of states la) and Ib), there is some value of w between -00 and +00 which satisfies the condition tlW :: E a - Eb' (Note, if we had the integral from 0 to 00, we should have zero from states for which En - Eb was negative. It is for this reason that we let the integral range from -00 to +00.) Thus, changing the variable of integralion from w to liw, we get 1 j j(w)dw = 2 n ~:::
Jo
~
I
L
(mlm2 m J ···lp;l m lm2 m J .•• )
(3.25)
ml,ml,m:s, ..
Now, since pz::
Ej Pzj,
pi:: L,pzjpzk
(3.26)
.
j,k
There are two types of terms: j '" k and j :: k. We examine the first kind first. Let us consider j :: 1, k:: 2. Then, holding m2' m3, m4 ... fixed, we can first sum over mt. Now,
so that summing over m 1 gives us
[~(mIIPh;lml)] (m2Ip2zl m 2)
(3.23)
n,b
DPIAIP')(P'IBIP") = (PIABIP")
(3.28)
Now, Eml (mllplzlml) :: O. This may be seen by noting that when we lake ml as eigenvalues of liz, all the diagonal elements of liz and Plz are zero. Or, alternatively, one may let ml be the eigenvalues of liz' But for every +m value there is a corresponding negative one, so that
ml
DUl it is a basic Iheorem of quantum mechanics that for any complete set and for any operators A and B,
72
where the symbol "TT" stands for "trace" or sum of the diagonal matrix elements. Another imponant theorem tells us that when we go from one complete set of onhogonal functions 1,8) to an alternative one I() [1,8) can thus be expressed as a linear combination of the IO's], the trace is unchanged. We may therefore choose any complete set of functions to compute the trace. In fact we shall choose a set of functions that is simply the product of the individual spin functions of quantum numbers mt> m2, m3 ... mN for the N spins. Therefore
L,(mllplzlmd:: "(ft L,(mI11lzlml):: 0
~
w
(3.24)
•
-~
I+~
o
1
L(al";la) = 2h T, (";)
f(w)dw:: 2h
Therefore we can compute either (&.;2) directly or compule it from the calculations of (w 2 ) and (w). (We shall do the laneL) .
1
j j(w)dw = 2h 0-
j j(w)dw
(&.,') = (w') _ (w)'
~
.
(3.29)
ml
Therefore the contribulion from tenns j :/: k vanishes. For j :: k, we get, taking j :: I,
1,8')
(3.23,)
(3.30) 73
The matrix element is independent of m2, mJ, and so on, but it is repealed for each combination of the other quantum numbers. Since there are (21 + I) values of m2, (21 + I) values of m3, and so on. we get the matrix for each value of ml repeated (2I+I)N-I times. On the other hand. using Trl to mean a trace only over quantum numbers of spin 1. we have that Trl {llIz} = TTJ {Jll y }. This equation
To compute the average frequency or first moment rigorously,
=
wj(w)dw
(w) =
is most simply proved by first evaluating Tq {Plz}' using eigenfunctions of liz'
Then
+1
Tq {Plz} = "(2 112
L:
m2
(3.31)
In a similar way. Trl {ply} may be evaluated by using eigenfunctions of Ily: +I
L:
j(w)dw
o
m=-1
Trl {PIy} = ..,2h2
J -,=~,----- J
m2
(3.3Ia)
m=-I
Therefore Trl {PIz} =Trl {llI y } =Trl
tlltJ = iTrl {pi}
+=
J
wj(w)dw =0
because the integrand is an odd function of w. We therefore are forced to compute r.'wj(w)dw:
J=o
I
wj(w)dw = ",
.
f(w)dw = 21'l72h2 I(I; 1) N(2I + l)N
I = -,
+=
J
(alpzlbXblpz[aXhw),(E. - E, - hw)d(hw)
L
(all'z[bXblpzla)(E. - E,)
.
(3.34)
h E.>E. (B2)
o We tum now to a calculation of the effect of the dipolar coupling on the average frequency of absorption, (w). The existence of such a shift implies that the local field produced by the neighbors has a preferential orientation with respect to the applied field. Since such an effect must correspond to a Lorentz local field .t!J.H, it must be of general order XIIHo. where XII is the static nuclear susceptibility. X'I is given by the Langevin-Debye fomlUla: XII = N"(2h 2 I(I + 1)l3kT, where N is the number of nuclei per unit volume. If the distance between nearest neighbors is a. N ;:' Iht 3• we have, therefore, that t1H S! (,lda3)(-rliHolkT) S! Hloc(,IiHolkT). Since the nuclear Zeeman energy 7hHo is very small compared with kT. we see t1H is very small compared with the line breadth Hloc and is presumably negligible. Notice that the physical significance of our expression for t1H is that the neighbors have a slight preferential orientation parallel to the static field given by the exponent of the Boltzmann factor ("(IiHolkT). Hl oc has a nonzero average to this extent. Since few) of (3.17) corresponds to infinite temperature, it must lead to a .t!J.H = 0, and (w) = woo 74
L
o,b 0
Since there are N identical terms of j = 1.:, finally we get as our answer
j
(3.33)
-= (3.3Ib)
There are 21 + I diagonal matrix elements of PI, each of magnitude 72h2I(I + 1). Therefore
..,2h2 I(I + I) Trl {Jtlz} = 3 (21 + I)
k
is a bit more difficult than the calculation of f(w)dw. In (3.22) it was convenient to extend the limits of integration to go from -00 to +00. As a result. for every pair of energies E a and Eb' there was some frequency w such that Eo - E" = trw. regardless of whether E a was higher or lower than E". We cannot do the same thing for (w), since
The energies Ea and E" are the sum of dipolar and Zeeman contributions (--yhHoM + En), as we have remarked previously. We shall assume that the dipolar energy changes are always small compared with the changes in Zeeman energy and that the latter correspond to absorption near wo (our earlier discussion of the role of the tenns A. H, ... F shows us this fact). Therefore. since Eo > Eb. we write
Eo
= -"(TiHoM + En
Eb = -"(IiHo(M + 1) + En' Eo-Eo = flWo + En-Er./
(3.35)
By using these relations, we can write (3.34) as
=
J
I
wf(w)dw = 2"
o
L
(MalJ-lzIM + 10")
" '"I ,n,a' X (M + 100'1/lz1M a)(!iwo + EO' - EO',)
(3.36)
We shaJl first discuss the contribution of hwo tenn in the parentheses. It is 75
~o L: it
(MaIJJ;!:jM + la')(M + la'IJJ;!:IMa)
.
(3.37)
= EOI, / u Al'OI,PU,\I01 dr
/11,0',0"
= Eo:,(M'o'jPIMa)
Were it not for the restriction to M + 1, (3.37) could be converted to a trace by means of (3.23a). This restriction can be removed by using the properties of the raising and lowering operators and by noting that
.
(3.42)
Therefore
L
(MoIJJ-jM'a')(M'o'lp+jMo)(Ea - EOt')
/11 ,/If' ,CI',O"
•
L M,/I1',&,«'
Thus
(Moll1i~,~-lIM'o')(M'o'I~+lMa)
(3.43)
.T'{[1i~'~-)I'+) . (3.38)
Since Jl+ connects only states M ' and /11 in a matrix element (M' a'IJJ+JM (X) where M' = Ai + I, we can rewrite (3.37), summing over all values of M' as
A detailed evaluation of this trace shows that it vanishes. Therefore, combining the results of (3.36), (3.39), and (3.43), we get
f 00
wf(w)dw =
~: Tr {pi}
(3.44)
But from (3.24), OOJ
(3.39)
I j(w)dw = 2h T, {~~}
o Therefore 00
J
wj(w)dw
(w)
where we have used the facts that Tr {Jl~} = Tr fJl~} and
.
= / =
u~l'Ot'~uMO'dr
J
uA1'o,PEotlMOI dr
= EO'(M'a'IPIMa:)
.
Likewise, using the fact that 1{~ is Hennitian,
(M'O"I1t'~PIMa)
= / uA'f'OI,1t'3PI.lMO'dr
=
jerC!JU/lf'cr)· PUMO'dr
J
(3.45)
o
(3.40)
We have so far handled the hwo tenn of (3.36). The technique for handling the tenn E OI - E OI , is very simple. We know that }£~ IMa') = EO', 1M a'). Therefore, for an operator P, we have
(M'a'I~IMa)
=wo
j(w)dw
Tr {JJ;!:JJy -,lyp;!:} = "'(2 h 2 Tr {I;!:Iy - fIJI;!:} = ,2h2iTr {I:} =0
=~
(3.41)
The "average" value.of the frequency is therefore uns~ifted by me broadening as we had expected. To get the local field correction that we mentioned in our qualitative discussion, we should, in fact, have to go back to (3.17) and include the exponential factors that we deleted in going from (3.16). [That this is true follows from me fact that the expression fjH ~ Hloc("'(hHolkT) depends on temperalUre. The only place the temperature enters is in the exponentials.] We can compute the second moment (w 2 ) by similar techniques: 00
/ w 2 f(w)dw
~=~
J
O~
j(w)dw
o
Since we have already evaluated the denominator, all that remains is to compute the numerator:
76 77
00
Jo
2
w !(w)dw
1+00
Jw
:; 2
2
{Llw') =
j(w)dw
-00
1+00
J
=:2
-00
=
2'"
~7' ", I(l + I)~ L
4
Lw'(alp.lb)(blp.la)b(E. - E. - ""')dw
(3.47)
a,b
L...c E • - E.), (alp.lb)(blp.!a)
.,'
jl.:
is independent of j. There are then N equivalent sums, one for each value of j, giving us
3
L
{Llw') = -7' ", I(l + I) 4 I.:
(3.48)
= 'Hz + 'H'~ to get
where in Ihe "cross-Ierm" involving {[1tZ. 1Ja:] and l1t3, J.lz] we have used the basic relation true for any pair of operators A and B:
(3.50)
which is readily proved by applying (3.23a). If the dipolar coupling were zero only the first lenn on the right would survive. and of course the resonance would be a 6"-function at w::;; WOo In this case (w 2 ) = Therefore we sec that the first lenn must contribute w~ 10 (w 2 ). Explicit evaluation in fact verifies lhis result. The second. or '\,TOSS", leon vanishes, since every term involves factors such as Trl {Pb-}' The lasl term, when divided by j(w)dw gives
wfi.
k
~,'i ft2 1(1 + I) (~) L: (I N
- 3 c~s2 8jk )2
i,I.:
Now, by (3.21),
{Llw') = (w') _ {w)' Therefore, since (w)::: WO, we have
78
rjl.:
J
We can get a clearer understanding of (3.52) by considering an example in which all spins are located in equivalent positions, so that
I.:
By using the fact thai 'Hla) = Eala), we see, as in (3.42) and (3.43),
Tr {AD} = Tr{BA}
(3.52)
r~1.:
~ (I - 3 cos 2 6 j l.:)2 L.J ,.6
1 "
We can expand, using H
(I - 3 cos' 9j ,)'
N J,. I.:
(I - 3cos 2 6· )2
6
I'
(3.53)
Tjl.:
Each term is clearly of order (-1H{;,,/ where H~c is the contribution of the kth spin to the local field at spin j. The imponant point about (3.53) is that it gives a precise meaning to the concept of a local field, which enables one to compare a precisely defined meoretical quantity with experimental values. So far we have considered only the second moment for a case where all nuclei are identical. If more than one species is involved. we get a somewhat different answer. The basic difference is in the terms of type B in the dipolar coupling that connect states such as 1+ -) 10 J - +). If the two states are degenerate. as in the case when the spins are identical. B makes a first-order shift in the energy. On the other hand, when the states are nondegenerate, B merely produces second order energy shifts and gives rise to weak. otherwise forbidden transitions. It is therefore appropriate to omit B when the spins are unlike. 2 The interactions between like and unlike nuclear spins may be compared and me second moment readily obtained. If we use the symbol 1 for the species under observation. and S for the other species, the effective dipolar coupling between like nuclei is .. ,0
(rtJ)Il:::
I 2 2~(1-3coS26H)
,("h
L.
x,1
3 rk/
(3[zl.:lz1 ~ II.: ·1/)
(3.54)
In computing the second moment for like spins, the terms II.: • II do not can· tribute, since they commute with 1-'% [see (3.49)]. The coupling between unlike spins is
(3.51) 2 Van Vied: points out thlll omitting these terms for unlike spins lIS well as the terms C, V, E, and F for like sl,ins is crucial in computing (.1",,2). The !'i:lISOn is lhlll in computing (.6",,2), the ralher wellk satellite lines at "" 0 lind"" 2<.00 cocreipond to a typical f~uency fcom the center of the N50nancc, which is 110/1110<: larger than t1Kl1;C or the main transition. The second momenl mcUUI'(!S lile square or the rrequency deviation. Therefore, althoush the sal.ellitcs are down in intensity by (Jl loc / Jlo)', they contribute an amount quile comparable to lhe second moment. Since we arc concerned with the: width or the main lrllnsitiol'l, we do not wish to include the satellites. We must exclude the terms thal produce lhem fcom the Hamillonian.
=
=
79
(3.55) Equations (3.54) and (3.55) differ primarily in the numerical factor of the zzterm, {3.SS) being small by a factor of j. This numerical factor becomes ~ in
,
,
,,
,
...
the second moment, giving for the final answer.
(&.12)/5 =
~7h~h2S(S + 1)2. E (I - 3~s2 8jk)2 3
Nj,k
Fig.3.9. Unit cell or the ben'tcnc crystal. Solid lines represent molccutCll in the II = 0 plane; dashed lin~ represent moJccul~ b/2 abo,'c
, •
(3.56)
rjk
NOIice that it is S(S + I), nOI 1(1 + 1), Ihal comes inlo (3.56) expressing the fact that the local magnetic field seen by nuclei 1 is proportional to the magnetic moment 'YshJS(S + I) of the other species. The total second moment of the resonance line of spin I is given by adding the second-moment contributions of Hke nuclei 10 those of unlike nuclei.
3.4 Example of the Use of Second Moments Since the pioneering work of Pake and Gurowsky, numerous studies of second moments have been reported. A panicularly interesting example is provided by the work of Andrew and Eades (3.1) on solid benzene. By studying the various isotopic compositions in which protons were replaced by deuterons, mey were able to measure the proton-proton distance between adjacent protons in the ring and to show that at temperatures above about 90 K, the benzene molecules are relatively free to reorient about the axis perpendicular to the plane of the molecule. We shall describe their worle.. The three isotopic species studied by Andrew and Eades are shown in Fig. 3.8. The structure of the benzene crystal is very similar to that of a facecentered cubic crystal with me benzene molecules on the comers and face cenlers of the cube. However, although the sides of the unit cell are perpendicular, they are not equal in length, the a-, b--, and c-axes being, respectively, 7.44 A, 9.65 A, and 6.81 A. All benzenes have their planes parallel to the cystalline b--axis. A
rough sketch of the crystal structure is shown in Fig. 3.9, looking down the b--axis edge on to the plane of me molecules. The plane of Ihe molecules is represented by straight lines, solid for those atoms in the y = 0 plane, dashed for those b/2 above the y = 0 plane. (Since the samples studied by Andrew and Eades were polycrystalline, studies of the effect of the orientation of the magnetic field relative to the crystalline axes were not possible.) As we can see, there will be contributions to the second moment from nuclei within the same molecule and from nuclei outside the molecule. In principle, if one knew the location and orientation of all molecules, the only unknown parameter would be the distance R belween adjacenl protons in the ring. By using isotopic substitution, however, Andrew and Eades were able to obtain an experime1l1al division of the total second moment into contributions within and outside. We can see this readily by noting that replacement of a proton by a deuteron on any given site reduces the contribution of that site to the second moment by the factor 0': 0' =
~ i&lo(Io + I) 9 i~Ip(Ip + I)
(3.57)
where the subscripts P and 0 stand for the proton and the deuteron. By using the facts that 10 = l, Ip = (iDl2:lI") = 6.535 x 102 , (iPf17r) = 42.57 x 102, we have 0' = 0.0236. Thus, consider SI> the second-moment contribution from nuclei oUlSide the molecule. For C6H3D3, any given lattice position is equally likely to have a proton or a deuteron. Therefore the proton contribution to the second moment is cut by a factor of two. If all the lattice sites were occupied by deuterons, the second moment would be cut by the factor a, but since only one-half the sites are occupied by deuterons, the deuterons contribute O'Stl2. The total second moment 51, contributed by atoms outside the molecule is therefore
i,
+a)5'
5', = 51 2 + a51 2 = (1 2
FIg.3.8. Three spedes
60
or benzene studied by Andr~
and EtuJu
(3.58)
in which 0' is. of course, known. The analysis for the contribution from atoms within the molecule proceeds in a similar way. Let S2 and be the contribution for YiH6 and Y;H3D3.
52
81
D
Fig. 3.10.
__ -_'~~
~
\
\
I
,/R ~R
Posi~ions
of protons and deuterons and the rela·
(o'ig.J.ll. Second moment in Calla as a function of temperature
10
live distances
H I ' D , 2ll , \
"
.
I
~/
D ----
~.
H 9.7 gauss!
52
'-------r----------1.6gaU88 1
o.. L.L--*----.:;;--'--..:--::!:;-__ ~,__-,J :n 120 160 200
respectively. will be smaller than 52. since the nuclei in positions 2, 4, and 6 will give only Q times as big a contribution for the deuterated compound. By referring to Fig. 3.10 and recognizing the 1/r6 dependence of the contribution 10 the second mornen!, we see thaI
Temperature (X) II,
1)<>+\ ( 1+ m ~_, l+~+rh
-
.
(3.59) k
Thus we have for the second moments of 4H303 and C6H6. respectively,
S ""Sj +52=
C~Q)SI
i
+ 652 (3.60)
where a and 5 are known. Therefore, measurement of 5 and S' gives us 51 and 52 the separate contributions from outside and inside the molecule. The data for C&HsD provide an independent check. On the basis of such studies Andrew and Eades detennined the distance R between adjacent protons in the ring to be 2.495 ± 0.018 A, which is consistent with a prediction of 2.473 ±0.025 A based on the C-C spacing as detennined by x-rays, and an estimmed value for the C-H bond length. Of course one can combine the x-ray and resonance data to obtain the C-H bond distance. In principle. observation of the CI3 resonance would even pennit a detennination of the C-H bond length directly. TIle data we have mentioned were measured at temperatures below about 90 K. A second imponant result of Andrew and Eades was deduced by their studies of the temperature dependence of the second moment (Fig. 3.11). TIle rapid drop in second moment is due to the rotation of the benzene molecules about their hexad axis. Let us discuss this effect. The effect of rotation may be expressed very simply in tenns of the angles defined in Fig. 3.12. We consider a pair of nuclei j. k fixed in a molecule. the axis of rotation of the molecule making an angle with respect to the static field Ho.
e'
82
Fig.3.1Z. Angles important in describifl& the rotation of. molecule
Lei the radius vector from j to k make an angle 1jk with the rotation axis. Then, as the molecule rotates, the angle ejk (between Ho and the internuclear vector), which occurs in the factor I - 3cos2 gjJ. in the second moment, varies with time. Since the frequencies of rotation are high compared with the frequencies of interest in the resonan<:e)Jt is the time average of 1 - 3 cos 2 ejk that affects the second moment. Assuming the motion is over a potential well of threefold or higher symmetry, this average can be shown to be independent of the details of the motion, and
(3.61)
Equmion (3.61) shows that if the axis of rotation is parallel to the internuclear axis (1jk = 0). in which case the relative position of the two nuclei is unaffected by the rotation. the angular factor is unaffected by the rotation. On the other hand, if 1jk = 1r/2, 2 (I ~ 3cos 8jk )avg = -~(l - 3cos 2 g')
(3.62)
In a powder sample. we find all orientations of the crystal axes with respect to Ho. For a rigid lauice. we must therefore average (1 - 3cos 2 jk )2 over
e
83
the random crystal orientations. When motion sets in, we musl first average 1 _ 3 cos 2 0 Ok over the molion, to obtain the second momenl for a given crystal orientation. ~en we muSI average over crystal orientations. For interacting pairs, the cOnlribution 10 the second momenl of the rigid lattice {Llw 2 )RL Ihen goes as (3.63)
For the case of benzene, Andrew and Eades found Ihal the second moment of C6H6 from prolOns wilhin the molecule drops from 3.lOGauss 2 at low temperatures 10 0.77 ± 0.05 Gauss 2 at high temperatures. The assumption Ihat Ihe narrowing results from rotation about the hexad axis makes 'Yjk = 1r!2, since all protons lie in a plane perpendicular 10 the hexad axis, and predicts that the second moment should drop to 3.10/4 = 0.78 Gauss 2, in excellent agreement with the observed decrease.
where Ihe bar indicates an avemge over random orienlations of Ojk. When rOiation sets in, we have a second moment from the pair {Llw 2 )rot given by {Llw 2 )rot ex ({I - 3cos 2 Bjk )avg)2
(3.64)
where the "avg" indicates an average over rotalion, and the bar indicates an average of the orientation of Ihe rotation axis with respeci 10 Ro. By using (3.61), we get 2 {Llw )rot ex(l-
3cos 2 0')2 (
2 3COS ')"k J
2
-I)'
(3.65)
Since the cryslal axes are randomly oriented with respect to No, so are lhe rotation axes, specified by 0'. As a result
,
2 ,(3COS 'Yjk -
{Llw )rot = {Llw )RL
2
1))'
(3.66)
If ijk = 1r/2 (a pair rotating about an axis perpendicular to the internuclear axis),
the contribution of the pair interaction to the second moment is reduced by a factor of 4. 3 3 The justification for averaging 1 - 3 cos 2 (Jjk over the motion before squaring rather than averaging (1- 3cos 2 (Jjk)2 may be seen also by referring to the exact expression for X"(w), which was proporlional lo
0,'
The stales la) and 1fT) may be considered to involve both spin and rolational quantum numbers. But, since Ea - Eb is chosen to be ncar the Larmor frequency, lhe slales In) and 1fT) must have the same rotalional quantum numbers. Therefore, in computing the second moment, lhe lraee will be over spin variables only, but lhe angular factor will be a diagonal matrix clement in the "laUice" coordinales. But this means that we replace the classical 1- 3cos2 (Jjk by f 1l1.(1 - 3cos2 (Jjk)uLdT, where tiL is a lattice (in lhis case, rOlalion) stale. This procedure amounts to "averaging" 1 - 3cos2 (Jjk over the motion prior to squaring.
84
85
4. Magnetic Interactions of Nnclei with Electrons
4.1 Introduction So far we have ignored (he fact Ihal the nuclei are surrounded by electrons with which they can interact. In this chapter we shall consider the magnetic interactions. postponing until later the consideration of the strong eleclfOstatic effects that may be found when a nucleus possesses an electrical quadrupole moment. TIle magnetic coupling of the electrons to the nucleus arises from magnetic fields originaling either from the motion of the electrical charges or from the magnetic moment associated with the electron spin. The fonner gives rise to the so-called chemical shifts; the latter, to the Knight shifts in metals and 10 a coupling between nuclear spins. Both the chemical shifts and the Knight shifls have certain features in common. The total Hamiltonian of the electrons and the nuclei may be written as a sum of four teons: 'Hnz(H) + 'He(O) + HeZ(H) + Hen
where 'Hnz is the nuclear Zeeman coupling in the applied field H; He(O) is the Hamillonian of the electrons (orbital and spin) in the absence of H; 'Hez(H) is the electron Zeeman energy; and 'Hen is the interaction between the nuclear spins and the electron orbital and spin coordinates. If 'hen were zero, the nuclear spin system would be decoupled from the electrons, and the nuclear energy levels would be solely the Zeeman levels in the applied field H. The tenn 'hen corresponds to the extra magnetic fields the nuclei experience owing to the electrons. In a diamagnetic or paramagnetic substance. the average field a nucleus experiences owing to the electrons vanishes when H vanishes. However, since the interaction 'hez(H) polarizes the electron system, the effect of the electron-nuclear coupling Hen is no longer zero. We may say that the nuclei experience both a direct interaction with H through Hnz(H) and an indirect one through the interplay of 'hez(H) and 'Hen. The problem is very similar to the calculation of the electric field in a dielectric, in which we must add 10 the applied electric field the field arising from induced dipole moments in the other atoms. Systems such as ferromagnets possess electronic magnetization even with H = O. For them, the contribution of 'Hen is nonzero even without an applied field. 87
We shall consider the orbital effects first, starting with a review of the major facts about chemical shifts.
4.3 Quenching of Orbital Motion Classical elcctricity and magnetism tell us that a charge q moving with velocity v produces a magnctic field H at a point r ' away, given by
4.2 Experimental Facts About Chemical Shifts
q v x r
H= - - -' c
The most famous and most quoted example of chemical shifts is ethyl alcohol, CH;jCH20H (see references on "Chemical Shifls" in the Bibliography). The proton resonance consists of three lines whose intensities are in tile rulios 3: 2: I. If one possesses a highly homogeneolls magnet, each of these lines is found to possess structure that (as we shall see) is due to effects of eleclIon spin. The three lines are clearly due to the three "types" of protons, three in the CH3 group, two in the CHz group, and one in the DB. Evidently the nuclei experience fields of local origin that are different for different molecular surroundings. A comparison of the spacing in magnetic field between the lines as a function of the frequency of the resonance apparatus shows that the splitting is proponionalto the frequency. If we attribute the splitting to the fact that the nuclei must see a magnetic field L!JH in addition to the applied field Ho, we may say the resonance frequency w obeys the equation w = ,(H0 + L!J1I)
(4.1)
(4.3)
1',3
If we choose rather to ask for the field at the origin of a set of coordinates due to a charge at position r, then r ' = -r and (4.3) becomes
H=
1.. r
x v
c
1'3
=..J.....
r x mv
me,.3
=..J.....!:... me
(4.2)
If u is positive, we must use a larger magnetic field to produce the resonance than would be necessary for the bare nucleus. Of course we never do experiments on a bare nucleus, so that what we measure are the differences in u associated with different molecular environments. For protons, the entire range of u's covers about one part in lOs. For fluorine atoms, Ilowever, the range is about six parts in 104 , two orders of magnitude larger. Because of the small size of the shifts, they are ordinarily studied in liquids where resonance lines are narrow. Since the shifts should in general depend upon the orientation of the molecule with respect to the static field, single crystal orientation studies are of interest (see Chap. 8). As we have remarked, the chemical shifts are due to the orbital motion of electrons. It is imponant 10 contrast the orbital motion in solids or molecules with that in free atoms. We shall turn to this subject next.
(4.5) where p is the Bohr magneton (10- 20 erg/Gauss). For fluorine the average value of I/r 3 for the 2]) electrons is (4.6)
where ao is the Bohr radius. In other words, (1jr3) corresponds to a typical distance of A; the magnetic fields, to about 600,oooGauss. Such enormous fields would completely dominate the laboratory field Ho for typical experiments, in contrast to the facts. (Of course, in atomic beam experiments such large couplings can be observed.) We must understand why the large fields of free atoms are not present in solids or molecules. The disappearance of these large fields is also closely associated with the fact that, in most substances, the alOms do not possess permanent electronic magnetic moments; that is. most substances are diamagnetic. The term quenching of orbital angular momemwn is often applied to describe the phenomenon. Let us see how it comes about, by studying a particularly simple example. We shall consider an alOm with one electron outside closed shells in a pstate. We shall neglect spin, for convenience, although later in the book we shall return to the effect of spin in order to understand the so-called g-shifts in electron spin resonance. The three degenerate p-functions may be wriuen in either of two ways:
t
x/(,)
88
(4.4)
where L is the angular momentum of the particle about the origin. Equation (4.4) has a quantum mechanical counterpart, as we shall discuss. We see immediately, however, that for s-states, H = 0 aI the position of a nucleus, since s-states have zero angular momentum, whereas H 4- 0 for p, d, and other states of nonzero angular momentum. The magnitude of H is of order
where L!JH ex Ho. We may therefore define a quantity u, which is independent of H, by the equation
L!JH = -uHo
1. 3
yf(r)
,
zf(r)
or
(4.7) 89
(x ~Y) fer)
'f(')
,
(x
:;;y) f(r)
(4.8)
where fer) is a spherically symmetric function. The three functions of (4.8) are eigenfunctions of L z , the z-component of angular momentum and the m-values being, from left to right, 1, 0, and -I. The wave functions of (4.7) arc simply lincar combinations of those of (4.8). As long as the atom is frce, either set of wave functions is equally good, but if a magnetic field is applied parallel to the z-dircction, the set of (4.8) must be chosen.
,
,,'
-'I
,
'"
(b)
,
)
~"
'"
,
.,
If now we surround the atom by a set of charges in the manner of Fig. 4. 1, and for the moment assume that no static magnetic field is present, the degeneracy is lifted. The proper eigenstates are then those of (4.7), since a symmetric potential such as that of Fig. 4.1 will have vanishing matrix elements between any pair of the functions of (4.7). On the other hand, the diagonal matrix elements will be different, since the state xf(r) concentrates the electron on the x-axis, near to the positive charges, whereas the state vf(r) concentrates the electron near the negative charge. Clearly, xf(r) will lie lowest in energy, yf(,.) will be highest, and zf(I') will be between (unshifted in fact in the first order). The resulting energy levels are shown in Fig. 4.2. The ground state xf(r) may be written as
~ [(X :/zYl f(r) + (x :;;y) f(r)]
(4.9)
a linear combination of the m ~ + I and m ~ -1 states. The states m ~ + I and m ~ ~l correspond to electron circulation in opposite directions about the z-axis. Since they occur with equal weighting in (4.9), we see that xf(r) corresponds to equal mixtures of the two senses of circulation, or to zero net circulation.
1
yf{r)
"
zf(r)
j
90
xf(r)
is
;, = "1
(x!'"ay -y!...) ax
(4.10)
Therefore, for any wave function Uo,
(OILzrO)~ Juo~(x~ -Y~)ltOdT which, since
tlO
(4.11)
is real, may be written as
(OILzIO)""~ Juo(x~ ~Y:x)UO(LT
Fig. 4.1. (a) Four ehargcs plaeoo near an atom. The atom is assumed to be at the origin and the charges to b
xf(r) =
We can make a more precise statement by computing (L z ), the expectation value of the z-component of angular momentum. For generality, we shall make our proof for any wave function whose spatial part is reaL The operator for L z
I<1g.4.2. Splitting of the three p-slates by the charges of Fig. 4.1a
(4.12)
Since all the quantities in the integral are real, (4.12) shows that (L z ) must be pure imaginary unless the integral vanishes. But the diagonal matrix elements of any Hennitian operator are real. Therefore the integral vanishes and (OIL,IO)=O
.
(4.13)
It is clear that this proof holds for any component of angular momentum. We say that when
(OIL. 10) = (OIL,IO) = (OIL, 10) =
°
(4.14)
the angular momentum is quenched. Under what circumstances will the angular momentum be quenched? Clearly, what is needed is the possibility of choosing the eigenfunctions as real. In the absence of a magnetic field (which means that no spins are allowed to act on the orbit!), the Hamiltonian is real. If, moreover, a state is nondegenerate, its eigenfunction is always real, since it is the solution of a real differential equation (apart from an arbitrary complex constant factor, which clearly docs not affect the expectation value). Therefore (OIL",IO) ~ 0 for such a state. We conclude that whenever the crystalline electric fields leave a state nondegenerate, the orbital angular momentum of that state is quenched. The physical bliSis of quenching is that the external charges exert torques on the eleclron orbit, causing the plane of the orbit to precess. When the plane has exactly turned over, the sense of circulation is reversed. Crudely stated, the eleclron path has been changed from lying in a plane to being much like the path of the string in a ball of twine. Of course applicalion of a magnetic field will change things. We can see intuitively that a magnetic field in the z-direction will cause one sense of circulation to be favored over the other. The wave functions will readjust so that the ground state has a slight circulation in the favorable sense (the m ~ -I state will be favored). In temls of a small quantity e, this will make a new ground state: ~o =
I
[ ( X + i Y)
,ji (I -
e)
,ji f(,-j + (I + ,)
(X-i y ) ]
,ji f(,-j
(4.15) 91
As we can see, this change results from the admixture of a small amount of the state yl(r) into the ground state xl(I·). As we shall sec, the amount f: mixcd in is proportional to Ho, giving rise to a circulation that is proponiollal to Ho· We turn now to a closer look at the details of chemical shifts.
Chemical shifts arise because of the simultaneous interaction of a nucleus with an electron and that of the electron with the applied field Ho. A general theory has been given by Ramsey [4.1], but we shall present a somewhat different discussion, which breaks the calculation into two parts: I) the detennination of the electric currents produced in the molecule by the external field and 2) calculation of the magnetic field produced by these currents aI the nucleus. We shall work out the theory for one electron. We start by considering the Hamiltonian of the electron. To treat the magnetic fields, we must introduce two vector potentials, Ao and An, one associated with the magnetic field B o, the other with the magnetic field H n due to the nucleus. In tenns of Ao and An we have (4.16)
As is well known, there is more than one vector potential that will produce a given field. Thus, if H = 'il x A, a new vector potential A' = A + '\!¢ (where q, is any scalar function), will give the same field, since the curl of the gradient of any function vanishes. A transfonnation from A 10 A' is called a gallge transformation. We must be sure that the physical results of any calcul:uion are independent of the choice of gauge, that is, aT: guage-invariant. The effect of a magnetic field is introduced into the SchrOdinger equation by replacing the operator (Mi)'\! by (M)'\! - (qle)A, where q is the charge of the particle (q is positive or negative, depending on the sign of the charge of the panicle). The Hamiltonian then becomes
q)'
I ( p--A 1/=2m c
(4.17)
+V
where p == (li/i)'\!. If one uses n different gauge, AI = A + '\!¢(r), the new solution ¢' is related to the old one (t/J) by the (unitary) trunsfonn:uiOIl
",' = '" exp [ + (iqlt
In a shnilar manner the operator for angular momentum, r x mv, is
(4.18)
If we compure the expectation values of (li/i)'\!, (ljJ, Ul/i)\1ljJ), lllld (ljJl, (h/i)\1ljJ'), we find that they are not equal. Therefore, since any physical observable must be independent of the choice of gauge, we see that (fili)'\! cannot be the momentum operator mv. The operator for mv, rather, is (II/i)'\! - (qle)A, which is gaugeinvariant. That is,
The distinction between mv and pC=. (fdi)'il) is found in classical mechanics. Thus in tenns of the Lagrangian L, the definition of the canonical momentum pz is aLlai:, whereas the x-component of linear momentum is mi:. When a magnetic field is present, one finds pz =. mi: + [(qlc)A z ]· A quantity that will be of great importance to us is something which we shall call the current density jeT). It is defined as follows: j(r)
92
~ChefT1icl'Il Shifts"
in the Bibliography.
~..'L ':.("'.,,'" 2m I
",,,,,,') _ LA",'",
(4.20)
me We note that j(r) is a vector function of position, and il is real (Ihat is, has zero imaginary part). We recognize it as being q rimes the quanlum mechanical probability current. Explicit evaluation, first using'" and A and then using t/J' and A', shows that jeT) is gauge-invariant. Moreover, by assuming that t/J is a solution of SchrOdinger's time-dependent equation, one can show that
ae
d· . 0 lVJ+ fJt =
with
f!
== q¢"t/J
(4.21)
That is, j obeys the classical equation of continuity. For stationary stmes, ¢"t/J is independent of time and div j = O. j acts much as a classical current density. As we shall see shortly, such an interpretation is very useful in considering chemical shifts. The Hamiltonian for our electron acted on by two magnetic fields is therefore
l(q
'f)' +V
(4.22)
1/=- p--Ao--An 2m e e
where V representS all potential energy, including that due to fields that may quench the orbital angular momentum. II is convenient to define :l quantity 'If by the equlIlion 'If=p-!!.Ao (4.23.) c Since both p and Ao are Hennitian operators, so is 'If. Then (4.22) becomes 1
1(.=-2• m
2
q
([2
1ne
1ne
2
- - 2 ('If.A n +A n ·'If)+-2 2 An + V
(4.23b)
We shall choose An to be
A 1 See rderences under
G"-H
rx
4.4 Formal Theory of Chemical Shifts I
Ho = '\! x Ao
(4.19)
_ n -
J.1.XT r3
(4.24)
93
where ~ is the nuclear moment, since this vector pOlcnlial generates the field of a dipole. Since J.l is very small compared with electron moments, we expect to be able to treat it as an expansion parameter, and accordingly we drop the A~ term in comparison with me term linear in An. We thcn have
I , + V - --(:Jr' q ) 1i = -11" An + A n .:Jr 2m 2mc
(4.25)
.
In the absence of a nuclear coupling, (l/2m);r2 + V is simply the Hamiltonian of the electron in the presence of the stalic field. We shall treat the term involving An as a perturbation, computing the energy by using first-order perturbation theory.
lei us consider, then, the fim-order change in energy of a state whose wave function 'I/J is the exact solution of the problem of an electron acted on by the potential V and the static field Ho. The energy perturbation E1>crl is then
E......rt
'"
--q-
J
(4.26)
2mc
JAn.
I(1fw)"w + W·1fWldr
But, by using the definition of (4.20), we can write
2~ (~(",p)' +,p' ",pI
=
,r
(4.27)
given in (4.23) and of the current density in
-!L !:.w,,~
2m 1 = ;o(r) .
- ,p"~') -
,
!LAo,p'~ me
(4.28)
30(1') is the current density flowing when the static field is on. That is, 30 is
the current computed for the electron acted on by V and Ho (but not by the nucleus). Therefore E pcrt "" - ;
JAn. 30(r)dr
(4.29)
[Parenthetically, this fonnula gives us a general expression for the change in cnergy cE resulting from a change in ficld associatcd with a change SA in vector potential, in tenns of thc current 3(r) prior to the change SA;
SE ""
-~
J6A·
3(r)dr]
(4.30)
IJ~xr
Epert "" -~
"" -Il'
.
----;J . 30(r)dT
[~J X~D(r)dT]
(4.32)
l'
where, as we have remarked, Il is really the operator 7'11, but 3D(r) is simply a vector function of position. II is important to bear in mind that 30(1') is independent of the gauge of AD, the vcctor potc",ial of the static ficld. Equation (4.32) is identical in form to the classical intcraction of a magnctic moment Il with a current dcnsity jo(r), since thc quamity in the square brackcts is thc ficld H due to the current. Thcre is a good deal of similarity to the expression for the magnctic momcnt M of the elec(tons;
M ""
r/J.(1r' An + An . 7f)r/J dr
y2me where the integration is over electron coordinates. (Actually, An is a funclion of the nuclear moment ~, which must ilself be considered an operator. Thus Epc:rt will be an operator as far as the nuclear spin is concerned. We simply add il into the nuclear spin Hamiltonian.) By using the fact that 1r is a Hennitian operator, we rewrite (4.26) as
Epcrt '" --q-
we get
de J xjO(r)dT l'
(4.33)
Equation (4.32) contains the facts of the chemical shift. If we knew 30(1'), we could compute the resultant field at the nucleus. We can see that there are really two parts to the theory of chemical shifts: 1) finding thc current density io(r); 2) computing the integral of (4.32) once 30(1') is known. The latter problem is entirely classical and immediately involves one in such things as multipole expansions. Thus the effect of currents on an atom distant from the nucleus in question can often be approximatcd by a magnetic dipole moment. Since in general the current Jo(r) flows as a result of the presence of the static field Ho, dctennination of io(r) from first principles involves the solution of the quantum mechanics problem of an electron actcd on by elcctrostatic potentials and by a static field. On the other hand, in some instances one can guess the spatial fonn of 30(1') and use measured magnetic susceptibilities to fix its magnitude, a technique that has been used to explain the chemical shifts of protons in various ring compounds such as benzene, in which the currents in the rings are computed to give agreement with experimental (or theoretical) magnetic moments. Alternatively, one can tum the problem around and use the measured chemical shifts to detennine infonnation about magnetic susceptibilities of atoms, molecules, or bonds. Moreover, we can see that in general the chemical shifts will be most sensitive to nearby currents because of the 1/,.3 factor in the integral of (4.32), unless nearby currentS are especially small. We shall see shortly that the small chcmical shifts of protons compared, for example, to fluorine atoms results from the fact that the currents near the protons are relatively very small. In any event, (4.32) and (4.33) give concise statementS of what a chemical shift or susceptibility experiment measures about the currents induced in a molecule by the external field.
If we now set /,xr
An ""--3-
(4.31)
r
94
95
Then
4.5 Computation of Current Density We tum now to computing io(r). To do so, we need the wave function,p, which describes the electron when acted on by both the electrostatic potentials and the static field. We have, then,
E" _1- (p _2. A o)2
1# =
H=
where
2m
c
"0 = "0 + L'''O''n n
which gives us for the current, • .1.*n.l. .1. '" *) '" hq * * Jo(r) = -!t(} 2. ('flO v 'flO - 'flO v,po + L.J - 2• (,po 'V,pn - 'l./Jn'V,pokno ml n m)
(4.34)
+V
(4.43)
11q +L 2 .
(4.35)
n
ml
(l/J~ 'Vl/Jo -,po 'V1/;;I)e~o - L Ao1/;ov1o me
(4.44)
The terril
By expanding the parentheses, we find p2 q q2 2 1i = 2m + V - 2mc(P ·Ao + Ao .p)+ 2mc2Ao
2 " .q("o'V"o - ,,0'V"0) " J(,') (4.36)
Let us assume that we know the wave functions that are the solution of the Hamiltonian 1io in the absence of the extcrnal field:
,>,
+V 1i01/;" = E n 1/;" (4.37) 2m Then we can look on the tenns in (4.36) that involve Ao llS pcrturbing the energies and wave functions. We shall compute perturbed wave functions so that we can compute the effect of the magnetic field on the current density. Of course Ao goes to zero when lIo vanishes, being given typically by 1io = -
AO = ~Ho x r
hqW'V" -
,,'V,,') - LAo"',,
,,(nl1ipcr tl°) .1,
./.' _ ./.
'fIO-'fIo+L.J
Eo
tl
.
Ell
'fin
Bu'
"0
(4.47)
(
'V"o
=
J(.)
('~nM+ (X~Y)'VJ(')
giving .() hq ( " 2 Jo r = 2m XJ - yt)! (,.) 2
=2mkxr!(r)
(4.48)
The current therefore flows in circles whose plane is perpendicular to the zaxis. If we define a velocity vCr) by the equation vCr) = io(r)
q"',,
(4.49)
where q1jJ*'I./J is the charge density, we find
(nIJipcrd O)
96
Ji
hq
Let us define
Eo
+ iY)
X
=
2mc
=
*
(4.46)
(4.40)
we need keep only those parts of the perturbation that are linear in Ho, or by referring to (4.36-38), we take q fipcrl = (po Ao + Ao' p) (4.41)
enO
*
hq
,0(r)=-2'("0'V,,0-,,0'V,,0) nu
(4.39)
2m. me we can compute io(r) correctly to tcnns linear inlIo as a first approximation. To do this, we need,p and 1/;* correct to terms linear in flo for that pan of io(I') in the parentheses, but for the last tenn, we can use for"" the unperturbed function 1fio. Since we always have
(4.45)
is the current that would now when lIo = O. When tile orbital angular momelllllfll is quenched .w that t/Jo is real, we see that J(r) = O. and the current density vanishes at all points ill the molecule in the absence of Ho. lt is the tenn J(r) that gives rise. however. to the magnetic fields at a nucleus originating in the bodily rotation of the molecule. that is, the so-called spin-rotation interactions that are obselVed in molecular beam experiments. It is instructive to compute io(I') for afree atom in the In = +1 j)-$tate. lIo being zero. io(r) then equals J(I'), so that
(4.38)
Although this involves a particular gauge, we can see that in any gnuge, Ao will be proportional to Ho. In the expression for the current io(r),
jo(r) =
m'
En
(4.42)
v(r) =
!:...
k x r
m (x 2 +y2)
(4.50)
97
which is tangent to a circle whose plane is perpendicular 1 1.(r)l- -h m +y2
(0
the z-axis, so that
./:£2
(4.51)
This gives a z angular momentum of (4.52)
in accordance with our semiclassical picture of the electron in an m "" + I state possessing one quantum of angular momentum. We see, therefore, the close relationship in this case of the current density, the "velocity", and our semiclassical picture of quantized orbits. When the states tPo and tPn may be taken as real (quenched orbital angular momentum), we have J(r) = 0, and
11 io(r) = 2 .q l)£nO -£;IO)(tPO'VtP" -ljJn'V.pO) Tnl
n
LAotP6 Inc
(4.53)
For (4.53) to be valid, it is actually necessary only that the ground state possess quenched orbital angular momentum, but for excited states, we have assumed that the real fonn of the wave functions has been chosen. Let us now proceed to look at some examples. We shall consider two cases, an ,,-state and a p-state. It will turn out that the chemical shirts for s-States are very small but that, for p-states, the effect of the magnetic field in unquenching the orbital angular momentum plays the dominant role, giving chemical shifts twO orders of magnitude larger than those typically found for ,,-states. To proceed, we must now choose a particular gauge for A o. It turns out, as we shall see, to be panicularly convenient to take
Ao
= !Ho
xr
= !Hok x
(4.54)
T
although an equally correct one would be Ao
= ,Ho x (r -
(4.55')
R)
where R is a constant vector, or
Ao z =0
,
Ao:>: = Hoy
,
Aoy =0
div Ao =0
--q-[Ao' p+ (,), Ao) + Ao .p] 2mc
98
h
,
~('J
·Ao) = 0
llperl
=-2~eHo·(r-R)XP
(4.59b)
(4.59c)
qf' = --HoL,(R)
2me where Lz(R) is the z-component of angular momentum about the point at R. The choice of gauge therefore specifies the point about which angular momentum is measured in the perturbation. It is, of course, most natural to choose R"" 0, corresponding 10 measurement of angular momentum about the nucleus, since in general the electronic wave functions are classified as linear combinations of .s, p, d (and so on) functions. When the electron orbit extends over several atoms, more than one force center enters the problem. The choice of the best gauge then becomes more complicated. A closely related problem in electron spin resonance involving the g-shift is discussed in Chap. 11. Let us now consider an .s-state. Then the wave function is spherically sym· metric: (4.60) ~.(r) = ~.(c) It is clear that, since
0
L,tP. = 0, (4.61)
.
Therefore tno is zero for all excited states, and the entire current io(r) comes from the last term of (4.44):
(4.56)
io(r) = _LAo!J;5 = --q-Hok x r!J;;(r) (4.62) me 2me The current therefore flows in circles centered on the z-axis. The direction is such as to produce a magnelic moment directed opposite to Ho so thm it produces a diamagnetic moment. We see that the current direction will also produce a field opposed to Ho at the nucleus (see Fig. 4.3). It is interesting to n()(e that there is a current flowing in the s-state. There must certainly be an associated angular momentum, yet we customarily think of
(4.57)
where (p' Ao) means p acts solely on Ao. But since p = Ul/i)'V. (p·Ao) =
qh -2-HoLz me Had we chosen the gauge of (4.55a), we would have
1{pert =
(4.55b)
Then we have, from (4.41),
,.-
--'-(Ho x r)'p 2me (4.59,) = --'-Ho' (r x p) 2me We recognize that r x p is the operator for angular momentum in the absence of Ro. It is convenient in computing matrix elements to use the dimensionless operator (l/i)r x 'V for angular momentum. Denoting this by the symbol L, we can write (4.59a) alternatively as 1{pert =
(nl1tpertl¢,) =
In tenns of the Ao, (4.54), we have
1{......rt =
Then, using (4.54), we have
(4.58)
2
2
99
I
FigA.3. Diamagnetic curnlllt flow in an ,,-state atom, and the magnelic fields produced by the current
y
/I,
y
-q
/I
-I/-k
/I
+q
.,
1
yf(r)
"j
zf(r) xf(r)
+q
-q
s-stales as having zero angular momentum. We are confronted with the paradox: If s-states have zero angular momentum, how can there be electronic angular momentum in a first-order perturbation treatment if the first-order pcourbalion uses the unperturbed wave function? The answer is that the angular momentum operator has changed from r x (Ii/i)\! in the absence of a field to r x [(llIi)\7 (q/c)A] when the field is present. By using the changed operator, the unchanged ,s-state has acquired angular momentum. The angular momentum is imparted to the electron by the electric field associated with turning on the magnetic field, since this electric field produces a torque about the nucleus. There is a corresponding back reaction on the magnet. We note that since A is continuously variable, we can make the angular momentum continuously variable. By using typical numbers for H o and r, one finds the angular momentum much smaller than h. Does this fact violate the idea that angular momentum changes occur in units of II? No, it does not, since the electron is not free but rather is coupled to the magnet. The complete system of magnet plus electron can only change angular momentum by h, but the division of angular momentum between the parts of a coupled system does not have to be in integral units of Ii. We turn now to a ]>-state Xf(I') acted on by the crystalline field such as that discussed in Section 4.3. We duplicate the figures for the reader's convenience (F;g.4.4). The energy levels are then as in Fig.4.5. Let us consider H o to lie along the z-direction. In contrast to the s-state, the p-state has nonvanishing matrix elements to the excited states, corresponding to the tendency of the static field to unquench the angular momentum. For this orientation of Ho, the matrix element to Zf(l-) vanishes. That to yf(r) is (nl1i pert 10) =
J (x'!!'-oy H0 "J ' dr
--'-Ho~ 2mc 1
= +-2 q me
7
,
yf(I')
y..£..-)Xf(1")dT
ox
[yf(')1
(4.63)
iqflHo =---2me where we have used the fact that the function yf(r) is normalized. By using (4.63), we find GnO=
100
(nl1i per dO)
Eo
E"
. qhHo I
=1---
2mc Ll
(4.64)
Fig.4.4. Crystalline field due to charges at:l: =±a, .'1= z =0; -q at y = ± a, :I: = Z = 0
+q
The tenn 1/;0 V't/J"
t/Jo \71/;n
-
- t/Jn \71/;0
t/J1l V't/Jo =
.
of (4.53) is readily shown to be
(xj - yi)f2(,,)
(4.65)
It is conventional that the pMt of (4.44) associated with the excited stllles is called the paramagnetic current jp, since (as we shall see) it contribllles a paramagnetic magnetic moment. We call the last tenn of (4.44) the diamagnetic current jo. Then, using (4.44), (4.64) and (4.65), we get for our example 1i2 jp;= 2m
2
~c
Ho
Ll k x r f2(r)
(4.66)
and by using (4.44) and (4.54), q2 I 3D = - - -Hok x rlt/J1 2
me2
"
= - - 2 Ho(k
me
x r)x2 j2(r)
(4.67)
It is clear that both jp and jn flow in concentric circles but in opposite directions. However, although div jp = 0, the same is not true of jl). Since div j = 0 for a stationary state (j = jp + jo), there is clearly a discrepancy. It may be tmced to the fact that the wave functions used to derive jp and jo are not exact solutions of the crystalline field, but are, rather, only zero-order functions. The charges that give rise to the crystalline splitting will also distort the functions. For example, xf(,·), which points towards the positive charges, will presulllllbly be elongated somewhat, whereas yf(r) will be somewhat compressed. This will result in current flow with a radial component that will, so to speak, supply the circular currents of the dilllnllgnetic tenn. However, radial currents will affect neither the chemical shift (since they produce zero field at the origin) nor the atomic magnetic moment. Therefore we shall not search for better starting functions. The division between diamagnetic and paramagnetic currents would change if we had chosen a different gauge for Ao. However, if our solution for jo(r) were exact (to order No), the towl current jo(r) would be gauge-invariant. It is for this reason that our expression is so useful, since it holds regardless of 101
the gauge of Ao used to com pUle the currents or of lhe division of io between paramagnetic and diamagnetic tenns. It is imponanl to compare the relative magnitudes of jp llnd in- From (4.66) and (4.67) we get . __ . m x 2 .1 = _ . Ll 3D Jr ,~t.2 JP(r'f') /1 Inx
(4.69)
8
where x is measured in angstroms and .1 in electron volts. Thus, if .1 = 8eV (a fairly typical .1 for chemical shift problems), we see that ip is larger for x less than I A but that io is larger outside. As we shall see, the distance that is most important for typical chemical shifts is about 0.25 A, so the paramagnetic current dominates; however, for computing magnetic moments, a distance of I A or greater is more typical, muking il difficult 10 assess which factor is the more important. We can now compute the chemical shift fields Hp and HO duc to jp and
HI' =.!.jrx c ,,3 2
= 1i. 2m
jp
dr
L 2 Ho j
r
X
(k x r) f 2 (1')dr
(4.70)
.1 r3 Direct evaluation shows that the x- and y-components of FIp vlll1ish, leaving only the z-component:
2mc2
mc
Now, for any wave function t/J(,'), Ihe mean valuc of 1/1,3 is given by
) j ,I,I,pI'dT =
,
(4.72)
so that we see
j
x2 f2(r)dr = ( I ) = 1,3
1'3
j
y2 f2(r)dr
(4.73)
,.3
Therefore we find
H
- kh
p-
"q
-Ho ( I )
~mc2Ll 1'3
(4.74)
We note that HI' aids Ihe static field and is in fact proportional, in keeping with the experimental data. 102
j T x (k1'3 x r) x2 f2(1')
(4.75)
j
k q2 (x 2 + y2) 2 2 Ho"" - - 2 zHo 3 x f (7") (4.76) me r It is most convenient to average H o over all orientations of Ho with respect to the X-, y-, and z-axes. This can be shown to be equivalent 10 averaging HI) for Ho parallel 10 the X-, y-, and z-axes, in tum giving
I q' j(x2+y2)+(X2+z2)+(y2+z2)] 2 Ho = --3 - 2zHo 3 x j(l') me r =
-LHO(~) 3mc2 r
(4.77)
Since Ho = -oDBO. where (10 is the diamagnetic contribution to the chemical shielding parameter (j, we have uD =
3~c2 (~ )
(4.78)
an expression first derived by Lamb [4.2] to describe the shielding of dosed atomic shells. We can likewise average HI'; however, we note here that I-Ip is zero when Ho is parallel to the x-axis, since the penurbation gives zero when acting on the cylindrically symmetric function xf(1'). It is also convenient to assume that yf(1') and zf(1') are degenerate, both a distance .1 above xf(r), since this corresponds to the typical case of a chemical bond. Then we have
(4.71)
C~
-~Ho
Ho ==
turns 'out also to be in the z-direction only: (4.68)
where h 2/mx 2 has the units of energy (comparable to the kinetic energy of an electron whose de Broglie wavelength is x). By substituting numbers, we find . . x 2L1 JD = - J p -
The diamagnetic field. which is given by
Hp = and
0"1',
~
3
li
2
m
L. Ho ( I ) mc L1 1'3 2
(4.79)
the paramagnetic contribution to u, is
(TI~ = - } :~ ~:2 ~ C\)
(4.80)
If we take .1 = 4.3eV, and (lh- 3 ) = 8.891(l~, where ao is the Bohr radius (values appropriale to the 21' electrons of fluorine, with the energy chosen to be appropriate for the F2 molecule), we find 0"1' = -20 x 10-'1. uo is typically 10- 5 • We see Ihat this value of Up is quite comparable to the changes in 0" observed for fluorine compounds, whereas 0"0 is much too small to account for the effects. It is dear also why the range of fluorine chemical shifts is so much larger than thai of protons. Physically, the large fluorine shifts come about because the magnetic field leads to an unquenching of the angular momentum. The smaller .1, the more effectively Ho can "unquench". 103
What can we say is the cause of the s-st3le shift? One simple picture is to nOle that an s-state is a radial standing wave. Since the magnelic force is transverse 10 the radial motion. it produces a slow rol3tion quite analogous to the manner in which the Corialis force causes lhe direclion of a Foucault pendulum 10 tum.
As we have seen, for any reasonable values of il, the paramagnetic shielding lenn will completely overwhelm the diamagnetic tenn. What can we say about M, Ihe electron contribution to the alomic magnetic moment?
M=~JrXjodT
(4.81)
H=!J r xio dr c ,3
(4.82)
2c We contrast this with the shielding field
Clearly, the 1/r J factor makes H relatively much more senSitive to currents close to the nucleus. In fact we can quickly convert our Cannulas for shielding to Cannulas for average susceptibility X by recognizing that only the radial averages differ. Thus the paramagnetic and diamagnetic currents contribute Xp and XD.
respectively. to the susceptibility X. where X=XP+XD
M=XHo
(4.83)
We find (averaging over all orientations) I h 2 q2 I I q2_ Xp =-- --XD =-- --,' 2 3 m mc .1 6 mc2 There are anisotropies that, for both Xp and XD' are a substantial fraction of the average value. If we compare Xp and XD, we have h2 r 2 Xo = -XP~ 2.1 = -Xp
-;:2 .1
2"8
(4.84)
where r is measured in angstroms and .1 in electron valls. In general we must expect (r 2)1/2 ...... tA.. and .1 . . . 8eV. Therefore it is clear that XD ':!' - Xp· We cannot decide which lenn is the larger without specific examination. NOIe in particular that the mere fact that jp is dominant in producing the chemical shift does fJ()1 mean that it will be the major factor in determining the atomic susceptibility. The susceptibility. since it depends on currents far from the origin. is much more strongly influenced by the diamagnetic currents than is the chemical shirl because the diamagnetic currents are the more prominent at large distances from the origin. It is important to bear in mind that our particular choice of gauge has made 1ipert dependent on the angular momentum about the origin. For an exaci solution, this gauge is no beller than any other gauge. However, we rarely deal with exact solutions. There may Ihen be some physical preference in choosing a gauge that puts the perturbation in tenos of angular momentum measured about 104
the most prominent force center in the problem. Since our wave functions will be approximate. the approximations will then at least be so primarily because of failure to acCount for the small crystalline potential rather than for the much large-r atomic (central) potential. When we are dealing with chemical shifts in molecules, we find il very hard to deal wilh bonds. since two force centers, one at each nucleus (for a pair bond), are important. The only simple approximation is to treat the aloms as virtually isolated. laking the excitation energies as being those of excited bonds, and including such effects as ionic character by using non normalized alomic wave functions. Pople [4.3] has discussed Ihis problem. using a technique due to London. His resulls can also be obtained by using perturbation theory. A similar problem arises when calculating g-shifts, and this is discussed in Chap. I I. When one has ring compounds such as benzene, the interatomic currents are important. Then one chooses a molecular force center; that is, a gauge in which L z is the angular momentum about an axis around which the molecule has roughly cylindrical symmetry. For benzene. this is the hexad axis. About this axis only a diamagnetic current results. This current, which flows all around lhe benzene ring, produces chemical shifts al the proton positions. Of course. since lhe diSlance of the protons from the ring is comparable to the radius of the ring. the chemical shielding field is not accurately given by replacing the ring by a dipole. On the other hand. when the currents j(r) are well localized in an atom or in a single bond, the dimensions of which are small compared with the distance to the nucleus under study, we can represent the effect of the currents by a magnetic dipole. When we average the result over the random orientations of a molecule in a liquid. we find that the shift vanishes unless the atomic susceptibilily is itself a function of the orientation of the magnetic field with respeci to the molecule. One important contribution to atomic currents, which vanishes in :l liquid, is the contribution of currents in closed shells on atoms other than thm containing the nucleus under study. The result in the liquid can be shown to follow simply because the current distribution in a closed shell is independenl of the orientation of Ho with respect to the molecular axes. We should emphasize that an attempt to compute the closed-shell contribution to shielding fields byapproximmc methods is dangerous, since one may find that the paramagnetic and diamagnetic COntributions are large. Their algebraic sum (which is zero for an exact calculation in a liquid) may be nonzero unless a very accurate computalion is made. It is therefore always safest 1) to judge the currenlS jo(r) on physical grounds, 2) always to choose gauges for each atomic current that puts ?-(p<,rl as proportional to the angular momentum about the most important atomic force center. and 3) to exclude from the calculation any currents that will give an exactly zero resull. Finally, we emphasize again the fact that the a priori judgment of whether a distant atom has a paramagnetic moment or a diamagnetic moment is not possible. but rather a detailed judgment of excitation energy and mean square atomic radius is essential. Moreover. for example, a paramagnetic moment on 105
another atom can produce either diamagnetic or pammagnetic shielding fields, depending on whether the anisotropic moment is largest when the inlcrnuclcar axis is perpendicular or paraUelto the direction of Ho· All the expressions given involve one eleclIOn only. If there are N elec· lIOns we generalize by adding subscripts "j = I to N" to the electron position coordinate. Thus we define AOj by the equation 'Vj x AOj = Ho
'Vj x Anj = H ..
(4.92)
(4.85)
Typically,
(4.93)
x.,.' Anj=~ ~
AOj = !Ho x(rj-R)
rj
(4.86)
'H..... rt = --q- "'(p .. AO' + Ao' .p.) .. 2mc ~ J J J J
where R is a convenient origin. We then define
(4.94)
)
(4.87)
"i J - -Ao' q J "'.J = -'V. C
so that the Hamiltonian, including both external magnetic field and the nuclear field, is
1"( E
(4.88)
q )' +V 7i:'7rj--A"j 2m j=1 c
If we define IJi to be the eltact solution of the N eleclIOn problem in the absence of the nuclear coupling, it obeys the equation ".' + ( _1 2m~ J
v)
'I' = E'I'
(4.89)
.
IJi is, of course, a function of the .,./s of all N electrons. Then we define the current associated with the jth electron as
"q . J[ -.(!P 2"11
•
'Vjtfl - tfI'VjW )
- -.!L.AOj!Pttf1] dTl ... dTj_ 1dTj+l'" dTN
Inc
(4.90)
where the integration leaves JOj a function of rj only. In temlS of (4.90), the nuclear coupling E pert is then Epert
=
-/.1-';I ~ Jr.J )
x Jo·(r·) /.: J 1 J
J
106
~ 2:J"'] x c . )
(4.91)
JOj(r)dT'
r~
(4.95)
J
J
Eltpressing JOj by means of (4.90), !P by means of (4.93) and (4.94), and taking divj AOj = 0, straightforward manipullllions give Ihe result: pdO) (01 Lj -fi-In) (nl LA", k
q2 h
H=,-,L me n +
dTj
The wave function !P, which is needed to compute the currents, is then found by using perturbation theory. Defining the functions tfl o, tfln to be solutions of the Hamiltonian 1tO in the absence of the external field, with eigenvalues Eo and
En'
To obtain eltplicit solutions, one must now assume reasonable N electron wave functions for !Po and wn . Ordinarily one will choose the functions to be products of one electron functions, or perhaps pair functions to represent a covalent bond. Although j labels electrons, one can often rewrite (4.91) so Ihat the sum over electron numbers is replaced by a sum over orbits, and in this way one can distinguish closed shell electrons from valence electrons. The formalism we have discussed is useful for obtaining a physical understanding of chemical shifts. The final result can be expressed more compactly in a single formula such as has been given by Ramsey [4.4]. To do this, we express the magnetic field, using (4.91) as H=
)
JOj(rj) =
Pj being (M)'V j'
1
{
En-Eo
(O'~AOk'Pk,n)(nlt'¥lo) ( En-Eo
q' - me' (01 Lrj
X
J
AOjIO)
(4.96)
)
If we further assume that AOk = ~Ho x rk and thai No = HOk, we get
107
, q2 h2 H=HOZm e n
"L:
+
J
\
E _~_
nUU
(01 p.dn)(nl tft[O)
I
u2(r)
En - Eo
_LHO(O[,,[k(xj+yj) _ 2mc
iXjzj _
2
Suppose we average 1{ over an s-state electron wave function 11(1'), as we would do to perfonn a first-order perturbation calculation of the coupling. There are a number of lenns 10 (4.98), similar to the tenns A, B, C, D. E, and F when computing lhe rigid lanice line breadth (see p.66). Let us pick out a tenn A, which depends on angle and distance as (I - 3 cos 2 U)/r 3. Then, apart from a multiplicative constalll, the average of such a tenn will be
(o[L:.f+!n)
~
r~
J
J
r~ J
6]IO)
iY
r~
J
~(l -
(4.97)
)
One can proceed to evaluate this expression directly rather than to compute the current density as an explicit function of position, as we did in our examples.
4.6 Electron Spin Interaction The coupling to electron spins produces effects using first-order perturbation theory when the electron spin moment is nonzero, as in paramagnetic or ferromagnetic materials. The Knight shifts (shift of tile resonance frequency in metals relative to their positions in insultors) are an example.' For diamagnetic materials, one must go to second-order perturbation theory to obtain nonvanishing spin couplings. One important class of phenomena that then results is the coupling of one nucleus with another via the electrons. These couplings give rise to fine struclure of resonances in liquids and to either narrowing or broadening of resonance lines in solids. For example, the indirecl couplings make the pure quadrupole resonance in indium metal be ahom ten times broader than Ihat computed from the direcl nuclear dipolar coupling alone. However, Ihere is no chemical shift associated with elcclron spin for diamagnetic substances. We shall discuss this poinl aI the end of Secl.4.9. We sian by discussing Ihe foml of the magnetic coupling between an electron and a nucleus. As long as the nuclear and electron moments It n and P e are far enough apan, we expect their interaction to be thm of a pair of magnetic dipoles, the Hamiltonian being 1{ = Me' Mn _ 3(l1c' r)(/t ll 1.3 1.5
•
r)
(4.98)
where r is the radius veClQr from the nucleus to the electron. As long as the electronic wave function is a I>-state, tiography.
108
2
2
3 cos 8)r
(4.99)
where dO is an element of solid angle. If we do the angular integral first, it vanishes, giving us a result of zero for (4.99). On the other hand, if we were to integrate first over r, we would encounter trouble near r 'Or 0, where u 2 (r) = u 2 (0) :;. 0, giving a logarithmic infinity. Since we can get either zero or infinity, depending on our method of calculation, it is clear that we cannot simply ignore the conlfibutions when ,. is small. From what we have said. it is evident that the dipole approximation has broken down. There are two effects thm come in and which have been neglected. First of all, we know the nucleus has a finite size. To the extent that the nucle:\r magnetic moment results from the bodily rotation of the nucleus, the currents are distributed over the nuclear volume. From the electron viewpoint the spin momellls of the nuclear particle are also spread over a comparable region, since the nuclear panicles effectively possess much higher frequencies of motion than does the electron (the nuclear energy levels are widely spaced in energy compared with those of electrons). A second effect is that the electronic coupling to the nucleus, when computed by using a relativistic theory (the Dirac equalion), shows a marked change when the electron is within a distance e 2 /mCl of the nucleus; e2 /mc 2 is the classical radius of the electron, ro, and is about 3 x IO~ 13 em. The electron is effectively smeared out over roo Since the radii of nuclei are given approximately by the fonnula r:= 1.5 x 1O- 3 A 1/ 3 cm
(4.100)
we see that the nuclear size is comparable to the electron radius roo Of course, completely aside from all these remarks, (he fact that the electronic potelllial energy is of order mc 2 near the nucleus shows us that a relmivistic theory is advisable. We shall first give a simple classical derivation of the interaction for s-states, and then we shall discuss brieny how the Dirac theory exhibits the s:l1ne featqres. The theorems on the relation between magnetic fields and CUITCnts, which we developed in discussing chemical shifts, will show us that our result is really rigorous for the contribution 10 Ihe coupling resulting from bodily motion of nuclear charges. Finally. since a volume distribution of magnetic moment (as produced by the spin momenls of the nuclear particles) is equivalenl to a current distribution, our result will also include the cOnlributions of inlrinsic spin. Thus, although simple, our calculation is in fact rigorous in the nonrelmivistic case. 109
We shall represent the nucleus by a charge q going in a circular path of radius a with velocity v. This. effectively, is a current loop of current (qlc)(l/T), where T is the period of the motion. We can express If z ~ the magnetic field in the z·direction, due to the nucleus, averaged over the electron orbital probability density lu(r)1 2 - as
- J Hz =
Hz(r)lu(r)1 2dr
(4.101)
where H z(r) is the field of the current loop. We shall lake z to be nannal to the loop. The olher components of H can be shown to vanish when averaged, since lu(r)1 2 is spherically symmetric for an s-statc. If we draw a sphere of radius a about the origin, we can express If :(J') by means of a scalar magnetic potential either for r < a or r > a. It is straightforward to show that the contribution from regions outside the sphere vanishes from the angular integrations. If we express the scalar potential inside the sphere as a Slllp of products of spherical harmonics with radial functions, all contributions except the first term (the term that corresponds to a unifonn field within r < a) vanish. We can evaluate this tenn simply, since it is the only one thm does not vanish at ,. = O. Therefore (4.101) may be rewritten as
-Hz = J" Heu (r)dr 2
(4.102)
o where He is the field at the center of the sphere. We may approximate (4.102) by recognizing that u(r) varies lillie over the nucleus as -
2
411" 3
(4.103)
Hz = Heu (0)3a
The field at the center of t!le loop is simply H=!lrxv_!l'!!""k
e
c
r3
- c
(4.104)
a2
But the magnetic moment IJ. n of the nucleus is i:rra 2, where i is Ihe "current", or JL =!l..!..:rra 2k=k
neT
Qav
2c
(4.105)
Thus H
It is convenient to express the coupling as a tenn in the Hamiltonian that will give this interaction. This is readily done by means of the Dirac S-function:
8,
J:i = -3JLe . Pn6(r)
where r is now Ihe position of the electron relative to the nucleus. It is convenient also to re-express (4.109) in temlS of the nuclear and electron spins I and S. For the electron we shall use a gyromagnetic ratio IC' which is positive. but for the nucleus. in is to have an algebraic significance, being either positive or negative. Then we have
(4.106)
e-~
By substituting into (4.103). we find
871" 2 kHz = 3IJ.nU (0)
(4.107)
IJ. n = "'In Ii I
IJ. c = -"'IerIS
8,
2
?1.=31c/r.r,/·S6(r)
8,
2
110
(4.111)
We notice that in (4.108), the radius of the nuclear orbit has dropped oul. Clearly, we should get the same answer for a volume distribution of circular currents. Moreover, since the smeared nuclear spin moment is equivalent to a volume distribution of current, we have also included the intrinsic spin of the nucleons if we use, for "'In and /, Ihe experimental values. Equation (4.111) is therefore quile general. We see also that if we may flot neglect the variations in u(,·) over the nucleus, the answer will be a bil different. Two isotopes that have different current distribUlions will then have couplings which are /lot simply in the ratio of the nuclear moments. This phenomenon is the source of the so-called hyperfine anomalies. The treatment of the imeraction by the Dirac equation is somewhat more involved. We shall sketch the important steps but leave the details to the reader. The Dirac Hamiltonian for an eleclron (charge -e) is
?1. =
-0:"
'(cp+eA) - pmc2 + V
(4.112)
where 0:" and p are 4 by 4 matrices, V is the electron potential energy, and A is the vector potential. We can express 0:" and p in lenns of the two by two Pauli matrices, (T. and the two by two identity matrix 1 as
(~ ~)
fi=
(~ ~,)
(4.113)
The wave functions 1Ji, which are solutions of (4.112), are represented by a column matrix of four functions, but these are most conveniently expressed in tenns of the functions !lit and IJiz, each of which is a column matrix wilh two elements:
The effective interaction energy E with an electron moment lJ. e is then E=-3IJ.c 'lJ. nU (0)
(4.110)
which give
a=
_ 2IJ.n
(4.109)
(4.114) (4.108)
The eigenvalues E of ?1. may be written as 111
E=E'+me 2 where E'
(4.115)
The tenn (4.12Ia) can be rewritten as
is the energy measured above me2 , so that for a free panicle at rest,
,'r"
2,. )' ( 21' + 1'0 (2mc 2)2
E'=O. If we define ll'=ep+eA
(4.116)
and define I/> as Vie, we have I (E + el/> + 2me2)!Pl + u - 1l'!P2 = 0
(4.117)
(E' + el/»!P2 + u - 1l'!P 1 = 0
where I/> = ell' is the potential due to the nucleus. As is well known, PI is much smaller than P2 in the nonrelativistic region. One customarily calls P2 the "large component". For s~states in hydrogen, P2 is also much larger than PI, even at the nucleus. One can eliminate !PI (still with no approximations) to obtain a Hamiltonian for rJi2, 'H' such that (4.118) By tedious manipulation one finds that
?-i'= EI
1
I/>
2
+e + me
2(e 2p 2+ e2A 2+2ecA_p_iecdivA
+ eflcu ° '\1 x A)+,
enc
'12
1 2encuo'\1xA E +e'f'+ me "2 1
and
(4.120a)
2
(E/+e
(4.12Ia)
For our problem, the nuclear coupling is given by introducing the vector potential A = JL n x r/r 3. Therefore \7 x A is simply the magnetic field of the nucleus computed by using the dipole approximation. As 10llg as el/>« 2mc 2, (4.120a) is exactly the same as (4.98) and goes as 1/1'3. If, however, T is so small that el/>_2mc2 , the answer is modified. By writing el/> = e'lh', multiplying (4.120a) by T on top and bOllom, using e 2 /mc2 = TO (the classical electron radius), and neglecting E', we get for (4.120a):
~)~Uo\7XA ( 21'+1'0 2mc
1.3
r)]
(4.12Ib)
The tenn in the square brackets goes as Ih.'l. For 1':»1'0, (4.12Ib) can be shown to be of order 1'0/1' times (4.120b); therefore, much smaller. However, when T :5 1'0, the radial dependence becomes less strong, going over to the hannless 111'0 near I' = O. The tenn is t/lerefore well behaved. It also has the feature that it does not average [Q zero over angle of s-states. It gives the answer of (4.108) for the magnetic interaction energy. We see that these two tenns are very similar to taking a S-function for sstates and asserting that a finite size of the electron prevents the radial catastrophe of the conventional dipolar coupling of (4.98). For computational convenience we may consider that the dipolar interaction of (4.98) should be multiplied by the function 21'/(21' + 1'0), to provide convergence. We shall now turn to the study of some of the important manifestations of the coupling between nuclei and electron spins, considering first the effects that are first order in the interaction and then effects that arise in second order. Further discussion of first-order effects will be found in Chap. lIon electron spin resonance.
lbe Knight shift is named after Professor Walter Knight, who first observed the phenomenon. What he found was that the resonance frequency of Cu 63 in metallic copper occurred at a frequency 0.23 percent higher than in diamagnetic CuCI, provided both resonances were perfonned at the same value of static field. Since this fractional shift is an order of magnitude larger than the chemical shifts among different diamagnetic compounds, it is reasonable to attribute it to an effect in the metal. Further studies revealed that the phenomenon was common to all metals, the principal experimental facts being four in number. By writing W m for the resonance frequency in the metal, Wd for the resonance frequency in a diamagnetic reference, all at a single value of static field, there is a frequency displacement .dw defined by Will
=Wd +.dw
(4.122)
The four facts are 1.
(4. 120b)
Now we no longer have an infinity from the radial integral in computing H:, and it is clear that the angular average makes (4.120b) to be zero. 112
x ("" x
(4.119)
where E is the electric field due to the nucleus. For our present needs, we focus on two tenns only:
'
1'3
4.7 Knight Shift'
(E +el/>+2me)
x (ieEoA +ieE·p - eu' E x p- eu·E x A)
u. [..':.
.dw is positive (exceptional cases have been found, but we ignore them for the moment).
3 See references to UNuclear Magnelic Resonance in Melals" in the Bil>liography
113
2.
If one varies Wd by choosing different values of static field, the fraclional shift .tJ.W/Wd is unaffected. The fractional shift is very nearly independent of temperature. The fraclional shift increases in geneml with increasing nuclear charge Z.
3. 4.
The fact that metals have a weak spin paramagnetism suggests that the shift may simply represent Ihe pulling of the magnetic flux lines into the piece of metal. However, the susceptibilities are too small (10- 6 cgs units/unit volume) to account for an effect of this size. As we shall see, however, the ordinary computation of internal fields in a solid that involves a spalial average of the local field is not what is wanted, since the nuclear moment occupies a very special place in the lattice - in fact a place at which the electron spends, so to speak, a large amount of time in response to the deep, attractive potential of the nuclear charge. As we shall see, the correct explanation of the Knight shift involves considering the field the nucleus experiences as a result of the interaction with conduction electrons through the s-state hyperfine coupling. If we think of the electrons in a metal as jumping rapidly from a10rn to atom, we see that a given nucleus experiences a m
necessary to specify the electron wave function, which is, of course, a formidable task from a rigorous viewpoint, one thai has not in fact been c:lITied Ollt because electrons couple to each other so strongly via the long range Coulomb interaction. Therefore we are forced 10 an approximation. We shall consider the electrons as being noninteracting - or at least only weakly so. Bohm and Pines have shown Ihal this approximation has considerable theoretical justification. For a review of the Bohm-Pines theory, see [4.5]. By means of a canonical transfonnation, they show that the principal effect of the Coulomb interaction is to give rise to a set of collective modes of oscillation, the plasma modes. The basic frequency of excitation of the plasma is so high that we may ordinarily consider the system to be in the ground plasma state. There still remain individual particle motions. The residual interaction between particles, however, is very weak and falls off nearly exponentially with distance. For low energy processes that do nOI excite the plasma modes, we may therefore treat the electrons as weakly interacting. We shall describe the system, therefore, wilh a Hamiltonian: (4.123) where He describes a group of weakly interacting eleclrons, H n is the nuclear Hamiltonian and includes the Zeeman energy of Ihe nuclei in the st:ltic field Ho as well as the magnetic dipolar coupling among the nuclei, and where 'Hen is the magnetic interaction between the nuclei and the electron spins. We omit the coupling of the nuclei to the electron orbilal motion because it gives effecls comparable to the chemical shifts. (Of course Ihe electrons in the metal are free so that the orbital effect is a bit different from that in an insulator.) It can be show~ that the conventional dipolar coupling between nuclear and electron spins, (4.98), contributes nothing in a cubic metaL For noncubic metals it gives rise to Knight shifts which depend on the orientation of Ho with respect to the crystalline axes. Since re.sonance in metals is usually perfonned on powders (to pennit adequate penetratIOn of the alternating field into the material, a problem that we may call the. "skin depth" problem), the anisotropy manifests itself through a line broadening. In the interest of simplicity, we shall confine our attention to the o-function coupling:
8.
114
(4.124)
j,l
where r/ is the radius vector 10 the position of the lth electron, and R· that to J the position of the jth nucleus. . Equation (4.124) can be rewriuen in a highly useful manner. When we diSCUSS electric charges in quantum mechanics we introduce a charge density operator e(op)(r) given by e(op)(r) '"
The dependence or the interaction on the relative orientation or nuclear and electron momenlS is mOSl readily seen by replacing the nucleus by a curre"t loop. 4
2",
Hen'" 3"Ye"Ynh LJlj' S{o(r/- Rj)
Lq/o(r/- r)
(4.125)
I
where q/ is Ihe charge of the fth panicle, and the sum goes over the particles. 115
In a similar way we introdoce electron magnetization WOp)(r) =
3
spin magnetization density operator of
L -1.1>5,o(r,- r)
,
(4.126)
The expectation value of this operator is the classical spin magnetization density. Using this definition, we get 14" = -
L
8; WOP)(R;H1"hlj )
(4.121)
1
Since we treat the nuclei and eleclTOns as only weakly interacting, we may write the complete wave function tP as a product of the (many panicle) wave functions tPc and VJn of the electrons and nuclei: (4.128) (Of course this wave function would be exact if ?ten were zero.) We shall then perfoon a perturbation calculation of the energy E ell :
E en =
J
l/J*?tcn!/JdTedTn
(4.129)
where dTc and dTn indicate integration over electron and nuclear coordinates (spatial and spin). Of course we shall wish to see the effect of (4.129) on lransi· tions of the nuclear system from one nuclear state!/JI! to another tPll'. Since the transitions are in the nuclear system. they leave the electron state tPe unchanged. In computing the energy of the nuclear transition £en - E en,. we have to compute both E en and Een ,. BOIh energies involve the integral over the electron coordinates. It is convenienl for us to postpone a specification of the nuclear states and to compute. therefore. the electronic integral
h~n
=
J!/J;1icn!/Jedrc
(4.130)
which is simply the first step in computing (4. I29) on the assumption of a product function. t/J = !/Je!/J1l or!/J' = tPe!/Jn'. We denote (4.130) by h~n to emphasize Ihal the nuclear coordinates still appear as operators. We can re-express (4.130) using (4.127) to obtain
, ,,8. hen = - ~ 3" M(R}). 'Yn hI}
(4.131)
1
where now M(R}) is the classical spin magnetization density at the posilion of the jth nucleus. The function !/Jc will itself be a simple product of one electron functions if we assume that the eleClTOns do not interact among themselves - or at least interact weakly. For the individual electrons we shall take the so-called Bloch functions. We remind ourselves of what these are: 1f the eleclTOns were thought of as moving in a one·dimensional box of length a (Fig. 4.6), the position coordinate 116
Ftg..4.6. Meh.t repr~lted by a box, with a polential deplh Vo
V, ' - -_ _--',_L.
being x, the wave functions would be sin kx or cos kx, whcre only those values of k are allowed that satisfy the proper boundary conditions aI x = 0 and x = a. In order to describe a situation in which a current can flow, it is customary to consider instead solutions exp(ikz), where now the allowed values of k are those that make the wave function the same at x "" a as it is at % "" O. The periodic boundary conditions for a three-dimensional box give solutions of the fonn !/J = e ik ' r
(4.132)
These solutions are modified in a very simplc way to tuke account of the fact that the real potential is very deep in the vicinity of the nuclei. The wave functions, called BlochfiUlctions, are then of the fonn
!/Jk
(4.133)
= uk(1')eik' r
That is, there is still a quantity k, the allowed values of which are given by requiring periodicity of the wave function on the walls of a box, but the plane wave exp (ik· 1') is multiplied by a modulating function uk(r), which is a function possessing the lattice periodicity. A typical uk(r) peaks up strongly near a nucleus. The fact that we explicitly label u with a subscripi k points out that U will in general vary with k. We shall need 10 add a spin coordinate as well, giving us finally a function ./.
"'k. = uke ik·
(4.134)
r.1.
"'.
where tP. is a spin function. The wave function for the N electrons, tPe, will then be a product of tPk. 's, properly anti symmetrized to take account of the Pauli exclusion principle. We can do this readily in terms of Ihe permutation opcrator P [4.61:
,p. = ~ L(-i)P p,p..(I),p.,.,(2),p", .,,(3) ... ,p••••(N) vN!
(4.135)
p
where the symbol (-I)P means to take a plus or minus sign, depending on whether or not the pennutation involves an even or an odd number of interchanges. The factor l/.,fiiff is, of course, simply noonaliz:ltion. Let us compute h'en)'. the contribution to (4.130) of the jth nuclear spin, and choose the origin of coordinates at that nuclear site (Rj = 0). Then we have
Ken} = 8; 7e'Ynh2 Ij'
J¢; L
S,6(1")¢edTe
(4.136)
I
Since the operator S I 6(1") involves only one electron, there are not contributions to (4.136) from teons in which electrons are exchanged, and therefore we get 117
r (4.137) We now assume the electrons are quantized along the z-dircction by the external stalic field Ho. Then the only contribution to (4.137) comes from 5 zl ' (We could alternatively have kept juS! that pari proportional 10 [zj, assuming the nuclear spins to be quantized along HQ. the result being the same.) We can write the results of (4.137) as 871"
-2
'""
2
(4.138)
31'cln h Izj L-!Uk,s(O)1 1»sP(k, s) k,8
where the sum is over all values of k. s, and where p(k,s) is a factor that is 1 if k, s are occupied by an electron, zero otherwise. The factor 1»8 is the m yulue of the state "l/Jk,s; hence it is +~ or -~, and of course llk,s(O) is the wave function evaluated at the position of nucleus j. If we average Ihis expression over a ScI of occupations P(k. s), which are representative of the temperature of the electrons, we can write for the effective interaction with jth nucleus,
(4.139)
where f(k, s) is the Fenni function. For the electrons at absolute zero, f(k, s) is I for all k, 5, which makes the total (that is, spin plus spatial) electron energy less than the Fenni energy EF and is zero for energies greater than E F • At temperatures above absolute zero, f(k,s) is modified within about kT of EF (Fig. 4.7). Of course f(E) is
1 1 + e
(4.140)
The notation f(k, s) means f(E), of course, where E is the energy of an electron with wave vector k and spin coordinate s. Typically, E = Ek+Espin, where Espin is the energy associated with the spin orientation and Ek is the sum of the kinetic and potential energies of an electron of wave vector k. (We shall call Ek the
f(E)f---_~
(aj
f(E)
'---=<-]l
(b)
i\ £'1,.
E
E j ,'
E
Fig. 4.7. (a) Fermi rundion feE) at absolute zero. (b) Fermi rundion at a temperature T above absolute zero 118
(4.141) As we can see, the quantity in the brackets is (apart from a minus sign, since IJ- e = ~l'enS) the average contribution of state k to the z-component of electron magnetization of the sample. We shaH denote it by I'z k. The towl zmagnetization of the electrons, /-lz, is then '
~=L/-lzk
(4.142)
.
k
If we define the total spin susceptibility of the electrons,
Ilz = X~Ho
x~,
by
(4.143)
and define a quantity
xi:
by
-/-lzk=Xk 'H0
81r 2" 2 31'c1'nf, Izj L.!Uk,s(O)! msf(k,s) k,s
f(E) =
translational energy of the electron). For example, Espin is the Zeeman energy of the electron spin in the static field Ho. There may be other contributions to Espin, however, from the electrostatic coupling between electrons, which depends on their'"'Telative spin orientation. If we consider a typical tenn in the sum of (4.139) corresponding to a single value of k, there are two values of m s , giving us
(4.144)
then (4.142) is equivalent to
X~ =
Lxi:
(4.145)
k
We can therefore write (4.141) as
- 8 1r l'n nIzj !u,.:eO) 12xi: H o 3 so that the total effective interaction for spin j is
8, H zj '" -31'n L. 1 Uk(O) I" XkHO
(4.146)
(4.147)
k
Our problem now is to evaluate the summation. It would be simple to make the assumption of completely free electrons, but it is actually no harder to treat the case of a material with a more complicated band structure. We shall do the laller, since the resulting expression will enable us to make some important comparisons between experiment and theory. We assume, therefore, that the energy of the eleclrons, apart from spin effects, is detennined by its k vector. For free electrons, the dependence would be
(4.148) so that in k-space, all electrons on a sphere of given k would have the same energy. That is, the points of constant energy fonn a surface, in this case a sphere 119
"
/
"
for any slates k having the same value of translational energy E1.:' Even when we allow the electrostatic coupling between the electrons to affect the energy to tum over a spin, it may be reasonable to assume Ihm this modification depends at most on Ek- We therefore assume that is a function only of the energy Ek: Xl(E k ) (4.151)
"
xl
'.
'. "(b)
(0)
xl::o
"/
We
Fig.4.8a,b. Intersection of two surfaces of constant energy wilh the k. = 0 plane. (:1) Circular section of a free elCdron. (b) A less syrnmetric seetion for a hypothetical "real" substance
(Fig. 4.8). In general the effecl of the lattice potential is to distort the surfaces 2 varies slowly as one from spheres. We shall assume that lhe function IUk(O)1 moves the point k in k-space from one allowed k-value to the next, so that we can define a density function to describe the number of allowed k-values in any region. Let us define 9(EI.:' A)dEkdA as the number of allowed k-values lying within a certain region of k-space, defined as follows: It is a small cylinder lying between the energy surfaces Ek and Ek + dEk (Fig. 4.9). Its surface area on the top or bottom surface is an element of dA of the constant energy surface. We denote the particular coordinates on the surface also by the symbol A in 9(E, A). The total number of states dN between Ek and Ek + dEk is found by summing the contributions over the entire surface:
therefore rewrite (4.150) as
Clln
L IllkCO)!2 Xt = JIUk(O)12Xs(E,,)g(Ek,A)dAdEk
•
If we have any funclion F of E, we define ils average value over the surface of constant translational energy Ek' (F(k)}e • as
xt
(4.149) We can use these functions to evaluate the summation by replacing it with an integral:
L
•
1",(0)1' xl, =
J
1".(0)I'xk9(E•• A)dE.dA
(4.150)
Now xl: depends on the Fermi functions f(k,~) and f(k,-~) :ll1d thus on the energy E1.:, and on the difference in energy of a spin in Slate k with spin up versus that with spin down ('YeftHo for free electrons). Therefore would be the same
xl
Fig.4.9. Volume in k-space associated with dEdA
JJ
g(E•• A)dA
J
F(k)y(E•• Jl)dA
J
(4.153)
1".(O)I'y(E•• A)dA = 9(E.)(I".(0)I') E,
(4.154)
J(1".(0)1')
(4.155)
giving
L
•
1".(O)I'xk =
E,X'(E.)g(E.)dE.
Now XS(E k ) is zero for all values of E k that are not rather ncar to the Fenni energy, since for small values of E k the two spin stales are 100 percent populated, whereas for large Ek' neither spin state is occupied. XS(EIJ must look much like Fig. 4.10. XS(Ek) will be nonzero for a region of width about kT around the Fenni energy EF. We are therefore justified in assuming (Itt~:
L
•
",,(O)I'xk = (1".(0)1') E,·
J
X'(E.)g(E,)dE.
(4.156)
The integral remaining in (4.156) is readily evaluated in teons of (4.145), since =
.'CE,JIL
LXt =
• J
J
xl:y(Ek' A)dEk dA
(4.157,)
x'(E.)9(E•• A)dE.dA
A
L+.''---_--:OEJ!
120
1 = -(E) 9.
ll1en we can set the integral ovcr dA in (4.152) equal to
=
.,
-,
F(k)9(E•• A)dA
(F(k» E, =
x~
"
(4.152)
Fig.4.10 Function XS(E",) versus E",
E k 121
which, performing the integral over A, becomes
x::::: JXS(Ek)e{E,,:>dEk
(4.157b)
Therefore, using (4.147), (4.156), and (4.157), we may say that the interaction with jth nuclear spin is
8.
,.]
(4.158)
-'Yn lllzj [ 3(1l1k(0)1 ) EpXeHo
This is entirely equivalent to the interaction with an extra magnetic field aH, which aids the applied field Ho, and is given in magnitude by the equation
aH
8'11"
2
(4.159)
We see that this formula has all the correct propenies to explain the exper· imental results. I) It predicts that a higher frequency is needed for the metal than in the diamagnetic reference. 2) The fractional shift is independent of w. 3) Since both (luk(0)12) Ep and are independent of temperature, all/Ho is as well. Since the larger·Z atoms will have a larger value of (llIk(0)1 2) Ep
r
corresponding to the pulling in of their wave function by the larger nuclear charge, the in<:rease of Knight shift with Z is explained. The Knight shift fonnula can be checked if one can measure independently aH/Ho, (Itlk(O)12)Ep and.t:. There is only one case for which all three quan· tilies are known (Li metal). The spin susceptibility has been measured by Schumacher J4.7] by a method we will describe shonly. Ryler [4.8} has measured (l"k(O)1 ) E by measuring the shift of the electron resonance by the nuclear moments. This shift,LlHe, is given by
8. (I Uk (0)1') EpXn • Ho :::: 3
(4 . 160)
IJ.H,
where X~ is the nuclear susceptibility of the U 7 nuclei. Denoling the number of nuclei per unit volume by N, we have ::::
N"(~1t2 I(I + I) 3kT
(4.161)
)L..
Thus, since X~ is known, mcasurement of LlHe gives (11l/.:(O)1 2 In order to enhance the size of the shift, RYler polarized the nuclei, using the so·callcd Overhauser effect (see Chap.7). It is necessary then to modify the formulas slighdy, but the principle remains the same. For comparison with experiment it is convenient to compute (lu/.:(O)1 2) E. using wave functions normalized to the atomic volume, which we shall cafl 122
X:.
Table 4.1 5
Ho = T(I"k(O)1 )E,X,
s Xn
PF. We shall use PA to denote the wave function density at the nucleus for a free atom. It is then convenient to discuss the ratio PF/PA for Li and Na. A comparison of RYler's values, theoretical values, and values deduced from combining SchwlIacher's measurement of X~ with measured Knight shifts is given in Table 4.1. There is excellent agreement among all three values. For Na, Kolm and Kjefdaas [4.9,10] find PF/PA = O.80±O.03. We can combine this value with the measured Knight shift to obtain a quasi-experimental value of ~. Before presenting these results, we shall describe Schumacher's direct measurement of
l\,/PA in Li 0.'19 ± 0.05 0.45 ±0.03 0.442±0.015
Kohn and KjetdafU lheordical Experimenlal ptus Knigh~ shift.) Ryter (experimenlal)
ex:
x:
The fundamental problem in measuring is how to distinguish it from the other contributions to the total susceptibility, which (in a melal) are quite comparable in size. The method employed by SCluwlOcher is to isolate the spin contribution by use of magnetic resonance. From the Kramers·Kronig relations we have for the electron spin susceptibility .t:: - 2 JX~dw x:-'. w
=
(4.162)
o
where X~ is the imaginary pan of the conduction electron spin susceptibility. For a sufficiently narrow resonance, we may neglect the variation in w across the absorption line, taking it out of the integrnl. We can then change the integration from one over frequency to one over field, using w:::: "(H: X• :::: -2 - I e lI" W o
= J "dw Xe
o
'J= "IH
1
= -2 -
'll"WO
Xe (
(4.163)
0
An absolute measurement of the area under the resonance curve will therefore enable one to determine X~. (Actually, the approximation of a narrow resonance is not well fulfilled in Schumacher's case; however, the final formula can be shown still to be correct. For a discussion of these points the reader is referred to Schumacher's paper [4.7]. We may also note that the resonant absorption of energy at the cyclotron frequency occurs degenerate with that of the spin. However, the rapid electron collisions broaden it so much as to render it unobservable. One may be confident thai it is only X~ that is being measured.) Absolute measurements of absorption are always very difficult. Schumacher circumvented them by making use of the nuclear resonance of {he Li7 or Na 2J 123
nuclei in the same sample for which he measured the conduction electron resonance. For the nuclei, onc has a spin susceptibility X~I' given by (4.161) lind for which X
s=~ n
00
"Ynj X"dH
1T WO
(4.164)
n
Ho = O. as with ferromagnets or anliferromagnets. lei us discuss the case of ferromagnetism briefly.~ Once again the electron-nuclear interaction will consist of the sum of the convellfional dipolar coupling 1td and the 8-state interaction 1ts : (4.166)
o
By choosing wo/2-;r So' 10 Me/s, Sc/llullacher could observe either the electron or nuclear resonances simply by changing HOI the remainder of the apparatus being left unchanged. The nuclear resonance occurred at abollt 10,0000au55, whereas the electron resonance was al only a few Gauss. Then, if we denmc the area under the electron or nuclear resonances by A c or An, respectively, we have
We shall average this over the electron wave function .,pe, as with the Knight shift, to get an effeclive nuclear Hamiltonian 1t~n' which will contain the nuclear spins as operators:
1t~n =
j .,p~(1td + 1ts).,pedre
X~ = 'Ye A c X~l 1'n An
(4.165)
Since X~I can be computed, we are thus able to determine X~. Note that we can measure the "area" in any units we wish (such as square centimeters on the face of an oscilloscope) as long as they are the same for bOlh resonances. We do not even need to know how much sample we have, since it is lhe same for both electrons and nuclei. The experimental values obtained are listed in Table 4.2 IOgether with various theoretical values. The first column of experimental and theoretical numbers [4.5] gives theoretical values based on non interacting electrons, blll an effective mass has been inttoduced to take account of the lattice potential. The effective masses are computed by Harvey Brooks, using the quantum defect method. The second column shows the theoretical values obtained by Sampson and Seitz, who took into account the electron-electron coupling by means of an interpolation formula of Wigner. The next column shows theoretical values due to Pines, based upon the Bohm-Pines collective description. Next we give values obtained by using the Knight shift and the Kohn-Kje/daas theoretical values of PFIPA . The last column shows Schumacher's results. (See [4.11] for subsequent resulls.)
where dre stands for integration over all electron coordinates (spin and spatial), and where the symbol G signifies that the spatial integration goes over the entire physical volume of the sample. By breaking G into the atomic cells G I, G2, ... GN of the N atoms of the crystal, we may fonnal1y interpret the contribution of 1td to (4.167) as arising from the summation of the dipolar fields of the electrons on the various atoms. The wave function .,pe will provide a detailed picture of the spatial distribution of the electron magnetization in each atomic cell. Evaluation of this tenn is identical to computing the local field due to a volume distribution of electron magnetization. Suppose the magnetization is unifonn throughout the sample. If the lattice has cubic symmetry, 1td will contribute an effective field given by the Lorentz local field:
4, -M -
3
_ a ·M
I,i Na
Sampson and Seitz
1.17
2.92 1.21
0.04
-
."j' WeSjo(r/- Rj)1/JedTe
8<, e h L H&j = - 3
Bohm-Pines
Knight shift. and theoretical Pp/PA
Schumacher
1.87 0.85
1.85±0.20 0.83 ± 0.03
2.08±O.1O 0.95± 0.10
If we took the wave function to be a product of one electron stales of quantum numbers fl, we would have, then,
8,
We point out that if Ryter' s value of PFIPA is used, the Knight shift value is raised, providing excellent agreement wilh that of Schumacher. The Knight shift calculation is closely related to the problem of the nuclear resonance in samples in which the electron magnetiz..1tion does not vanish when
(4.169)
I
H,j =
124
(4.168)
where M is the magnetic dipole moment per unit volume and 0' is a demagnetizing factor (in general a tensor) that expresses the effect of the "magnetic poles" on Ihe outer surface of the sample. For example ;; = 471i3(ii + jj + kk) for a sphere. The s-state lenn may be interpreted as contributing a magnetic field H&j at the jth nucleus.
Tabte 4.1 x: (all values are 106 cgs volume unit..s) Free eledrons
(4.167)
G
-310"
L:
(PIS,(r - Rj)IP)
111) of II
set
(4.170)
fJoccupied
~ See the references lisled under "Nuclear Hesonance in Ferromagnet.s and Antifcrromagncts" in the Bibliography.
125
"
where "occupied" means that we include in the sum only those states 1m containing an electron and we have omiued lhe subscript 1 from Sand r. By using the fact thai Ile = -"YenS, we have
8. H,j = 3"
'" L
(PI".'(r - Rj)IP)
(4.171)
lJoccupicd
Since the matrix element involves coordinates of only one electron, the various values of I now appear as Ihe values of p Ihal 3rc occupied. In a substance such as iron, we may think of some values of {3 as corresponding to closed shells, some to the 3d band, and some to the 48 band. We shall discuss these contriblllions shortly. The contribution of the tenn 7-{d is somewhat different for a ferromag-
net than for a paramagnet. For the latter, the magnetization is tlllifonn in both magnitude and direction for ellipsoidal samples, nnd the simple dcmugllclizing arguments follow. For a ferromagnet the magnetization within a domain is uniform, but the various domains have differing magnetization vectors. Thus, for a soft ferromagnet in zero applied field, the magnetization averaged over a volume large compared with the domain size is zero. The density of magnetic poles on the outer surface therefore vanishes. Within the body of the ferromagnet, div M "" 0 even at domain boundaries. If, then, we calculate the dipolar contribution to the magnetic field al a nucleus, we may proceed as follows. We draw a small sphere about the nucleus, of radius small enough 10 lie within one domain. We compute the field due to magnetization on atoms within the sphere by a direct sum. The atoms outside the sphere are treated in the continuum approximalion. For cubic symmetry, the atoms within the sphere give zero tOial contribution. The atoms outside the sphere contribute as a result of the surface pole density on the inner sphere and the Ollter sample surface. The former is the contribution 411" MI3, where M is the magnetization within the domain containing the nucleus. The latter contributes - ;; . M ' , where l\t! is the magnetization averaged over a volume large compared with ;\ domain size. TIle total field seen by the jth nucleus HTj is therefore given by
4.
HTj""Ho+
3
_ M-
0'
'M+Hsj
(4.172)
where lIo is an externally applied field. Although lIo and M ' vanish in zero applied field, HTj does nol. Therefore we have a "zero field" resonance. Such a resonance was first observed by Gos.tard and Portis [4.12] in the face-centered cubic fonn of cobalt. Using the Co59 resonance, the measured Hsj = 213,400Gauss. In iron, Hsj is 330,000 Gauss. Hsj has also been observed by means of the Mossbauer effect. It was discovered there that application of a static field H o lowered the resonance frequency, showing the fIsj points opposed to the magnetization M. The contribution from the 3d and 48 shells in iron is expected by Marshall [4.13] to give field of 100,000 (Q 200,OOOGauss parallel to the local magneti126
zation. Therefore the inner electrons must give a field of about 400,OOOGauss opposed to the local magnetization [4.14]. This phenomenon, called core polarization, was actually already known from e.lectron magnetic rcsonan.::e of paramagnetic ions for which the 48 electrons are missing. In principle the 3d electrons are incapable of giving an isotropic hyperfine coupling, since d-states vanish at the nucleus. However, the d-electrons are coupled to inner shell electrons electrostatically, the coupling for an inner electron of spin parallel to the d-electron spin being different from that of an electron whose spin is opposed 10 that of the d-electron. Consequently the spatial part of two wave functions such as the 38 are different for the two spin states. The spin magnetization of the two electrons does not add to zero at all points of the electron cloud. We can see from (4.171) that if the 3s electron densities at the nucleus differ, there will be a nonzero contribution from the 3s electrons to Hsj, even though their spins are opposed.
4.8 Single Crystal Spectra
"
We saw in Sect. 4.5 that the chemical shift for a p-state such as xf(1") depended on the orientation of the static field lIo, with respect to the bond axes. If Ho was either parallel or perpendicular to the bond direction, an extra field was induced, acting on the nucleus, which was parallel to Ho. For an arbitrary orientation, one can resolve Ho into X-, y-, and z-components. Ho z , Hoy, Ho z , in general gClling an interaction of extcrnal field with nuclcar spin involving bilinear products such as Hozly etx. Blocmbergen and Row/alld [4.15] showed theoretically and experimentally that the same thing is (nle for a Knight shift if one includes the conventional dipolar coupling (4.98) of the nucleus with the spins of the conduction electrons in addition to the Femli contact tcrm. Although the conventional dipolar tenn does not give rise to a net shift of the resonance if one averages over all orientations of Ho with respect to the crystal axes, as for a powder sample or as in a liquid metal, it nevertheless gives an orientation dependence to the Knight shift. To deal with these anisotropies, one expresses the bilinear form with coupling coefficients which are components of a tensor interaction. Then, writing the Knight shift J( and the chemical shift CT as tensors using a dyadic notation where
-
CT = iCTui + iuzyj + iCTzzk + jCTyzi + jCTyyj + jCTy:k
+ kCT;zzi + kCTzyj + kCTzzk and similarly for
J( ,
we get a Hamiltonian
-
'H = -'YliHo'( I - ';; + where
(4.173)
J(
)·1
(4.174)
I is the identity dyadic 127
2 2 ]1/2
=ii+ 11 + kk
+(1-uzz+Kzz) az
(4.175)
Explicit evaluation of u er /3 and I<':X/3 (a = x, y, z; both tensors are symmetric
fJ '"
x, y, z) shows that
Since the shift parameters ux X, Kxx, etc. are in general very small compared to I ,_and since 0').- +O'} +a~ = I, weean rewrite (4.183) to get
(4.176)
W = 'YHo[l + (Kxx - uxx)ak + (l{yy - Oyy )a~ + (Kz z - uzz)a~]
etc.
so that one can find principal axes. Denoting these axes by X, Y, Z with unit vectors i p , 1p ' and k p , we have
'; =ipuxxip+jpoyyjp+kpuzzkp
(4.177)
-
~
and similarly for K. In general, K and '; do not have to have the same principal axes, but we will assume for simplicity that they do. Then if Ho lies along a principal ax.is the resonance frequency is
= wo +waak +wba~ +wca~
'lip H = Ho +Hp
(4.186)
Then with Ho = kHo, where k is a unit vector along the laboratory z-direction, we get
(4.178)
Wz ='YHo(l-dzz+I
(4.185)
The approximation that the shifts are small could in fact be introduced in the original Hamiltonian. 1l1en we write 11. as a large part 1lo plus a penurbation
(4. 187a)
Ho = -jnHolz
wX='YHo(l-oxx+Kxx) :::::Wa+WO Wy = 'YHo(1 - oyy +Kyy) ::::: Wb +Wo
H p = +-yftHok· ( '(;
1<) ·f
(4. 187b)
But (4.187a) shows that I z commutes with the large pan of the Hamiltonian, so we keep only those tenns in H p which are diagonal in I z . Thus
where WO='YHo
(4.179)
If Ho lies in a more general direction we can express (4.174) by defining an effective field Heff
'H = -'YnHcrr' I
(4.184)
H p = 'YtIHo k · (;; Utilizing (4.183) we have k.ip=ax
(4.180)
where
(4.188)
(4.189)
giving us Herrx = (I - uxx + Kxx)Hox H effy = (1 - oyy + K)'y)Hoy
H p = jnHolz[ai-(ux
(4.181)
Herrz =(1- o z z +Kzz)Hoz
+ a~(J{zz - uzz)]Iz
(4.182a)
where
(4. I82b)
aX =sin Bcos tP ay = sin B sin tP
ax = HoxlHo
crz
(4.183)
ay = HoylHo az = HozlHo
= cos
(4.192)
B
The" 2 w=wo+wa sin Bcos 2 tP+Wb sin 2 B sin 2 ¢J+wccos 2 9
we get from (4.180) and (4.181)
126
(4.191)
It is often convenient to express the orientation of Flo in terms of spherical coordinates (B, tP), so that
Defining the three direction cosines aX, ay, and az by
[(I - ox X + K xx )2a~ + (1 -
(4.190)
H p = -jnHo[t + a~(J{xx - ux x) + a'~(l{yy - on')
Herr = JHJrrx +HJrry +HJrrz
W = 'YHo
KX x) + a}(uyy - I
+ cr~(uzz - I
0'
Then the resonance frequency w is given by W = 'YHeff
X -
O"y)'
+ Ky),)2a~
(4.193)
a"d
129
1i = -')'llHoP + (J(XX -
q, (4.194)
This resuh can be expressed in another manner which is useful when discussing the techniques of magic angle spinning which we take up in Chapter 8. Utilizing the trigonometric idenlities cos 2
q, = !(I + cos 21$)
2
4(1 - cos 21$)
sin 1$ =
(4.195)
,
(4.196) Defining (j
2
(4.197) = (
(LO for "longitudinal")
=:
(TR for "transverse") and correspondingly for the components of J(, we get iT) + (l( La -
- 11. = - '111Ho [ I + (/(
0 - 1)
+ ( I(TR ;
(4.198)
The two angular functions are linear combinations of the spherical hannonics Y2.no Therefore, they average to zero over a sphere. Since in a liquid there is generally rapid tumbling motion, such a spherical average is appropriate, giving W =wO+!(Wo+Wb+WC>
= '1 H o [
(I +
I<XX
4.9 Second-Order Spin Effects - Indirect Nuclear Coupling We have discussed the role of electron spin coupling to nuclei in paramagnetic or ferromagnetic materials. Since, in a diamagnetic substance, the tOlal spin of the electrons vanishes, Ihe nuclei experience zero coupling to the electron spins in first order. Effects are found if one considers the coupling in second oroer,e however. The coupling is manifested through an apparenl coupling of nuclei among themselves. the so.-called indirect coupling. The indirect coupling was discovered independently by Nahn and Maxwell (4.16] and by GUlOWSIey el al. {4.17]. The phenomena they observed are illustfUted In liquid PFJ , lhe by the case of PF3 , a molecule in which all nuclei have spin rapid tumbling narrows the line. It is found that both the pJI and F19 resonances consist of several lines, as illustrated in Fig. 4.11. Since all the fluorine nuclei are chemically equivalenl, the splittings cannot be due to chemical shifts. (FurthemlOre there is only one phosphorus atom per molecule but four phosphorus frequencies.) The fact that Ihe individual lines themselves lire narrow shows that the motion is sufficiently rapid to narrow the direct dipolar coupling. Moreover, the splittings are found to be independent both of temperature and static field. The number and relative intensity of lines are as though each nuclear species experienced a magnetic field proportional to the z-compooent of the tOlal spin of the other species. It was found that OWF = owp (see Fig.4.11), where OWf.' and owp are the f-requency separalions of adjacent lines in the phosphorus and fluorine spectra, respeclively. These facts indicated the coupling was somehow related to th~ nuclear magnelic momenls. The original explanation proposed was that one nucleus induced curreniS in the electron cloud, which then coupled to Ihe other nucleus. In a simple picture. the induced currents are represented by an induced electron magnetic moment.
4.
= !(
Often in working with solids one has a sample consisling of many small crystallites in the form of a powder. Then the orientation of the crystallites is random with respect to Ho. The anisotropic nature of the chemical and Knight shifts then gives rise to a line broadening and to specrra which have quite distinct sha~ depending on the relative values of woo Wh. and We' Such absorption spectra are called "powder patterns" and are discussed in Appendix I.
+ I<;y + I
+
--I
(4.199) For a system such as a bond xf(r) which possesses axial symmetry, it is convenient to take We to lie along the axis, so Ilial Wo = Wh. giving w=wo+ 130
'(
22 )+ (2wo-w.-w,)(3COS 0-1) ' (4.200) 3
3 w o+ Wb+ Wc
_
'w, J-
U
(b) L-
Fig. 4.11. (a) The p31 resonance in 1'10'3. The lines are equlIlly SPllCOO all amount &..,., and the intensilies are 1: J: J: 1. (b) The FU n'SOmmce in PF3 &
See rererfflccs to -Il"h Couplins" in the Bibliography. 131
If this moment were isotropic (as one changes the orientation of the molecule with respect 10 the nuclear moment), the coupling to a second nucleus would average to zero in a liquid, owing to the rapid random tumbling of the molecules. However, as we observed in connection with the chemical shifts, the induced moment is in general not isotropic. We can estimate the size of the coupling between the two nuclei from second-order perturbation theory. The first nucleus exerts a magnelic field I' l h(l/r 3 ) where (1/1'3) is the average of the inverse cube of the distance between the electron and the first nucleus that partially unquenches the orbital angular momentum, producing a fractional admixture of excited state of Il/e1i2(I/r3)/LJE, where LJE is the energy to the excited state. A complete unquenching would produce a magnetic field at the second nucleus of leMR3, where R is the distance between the nuclei (we treat the electron orbital magnetization as equivalent to a magnetic dipole). ll1erefore the order of magnitude of the nuclear-nuclear interaction energy EI2 on this model is El2 ~
I'IIe1i2(i"1r3) leh LJE
RJ 12 h
(4.201)
This formula fails by an order of magnitude or more in accounting for the facts. However, it has the virtue of making it seem reasonable that the splittings in PF J were an order of magnitude larger than those in PH3, since clearly this mechanism is closely related to chemical shifts that are always smaller for hydrogen than for fluorine. As was pointed out by Hahn and Maxwell, and GUlowsky and McCall, any mechanism such as we have described, and which will lead to a result that is bilinear in the two nuclear moments, must take a very simple fonn. Since the interaction is averaged over all molecular orientations, it can depend only on the relative orientation of the nuclei; hence it must be of the form (4.202) where Al2 is independent of temperature and field. These workers also pointed out that this panicular form would also explain the puzzling fact that, for example, there were apparently no splittings of f1uorines by fluorines in PF3 . We shall not give the proof here but physically the explanation is based on the idea that the interaction energy, (4.202), which depends on the relative orientation of spins, is unchanged if bOlh spins are rotated through the same angle. For equivalent nuclei such as the three f1uorines in PF J , one cannot rotate one fluorine spin without rotating the others by an equal amount, since the alternating and static fields are identical at all three f1uorines. Therefore the coupling between equivalent nuclei does not affect the resonance frequency. Ramsey and Purcell [4.18] proposed another mechanism utilizing the electron spins, which was substantially larger because, as we shall explain, it allowed the two nuclei to interact with nearby electrons, in contrast with the orbital mechanism where only one nucleus is on the same atom as the electron that is polarized. We may schematize their mechanism as shown in Fig. 4. 12. 132
AtomB
Atom A
I
I
Stat,r
t
Sta'" II
--------------
.I
Fig. 4.12. Bc<:ause of the bonds betwc<:n atoms A and B, the electron wave function is formed by an equal mixture of the state I, in which the electron Elloment on atom A points up (that 011 B, down), and state II, in which the spin orientation is reversed
In the absence of a nuclear moment the electron bond will consist of an equal mixture of the states I and II, shown in Fig.4.12. If now we put a nucleus on atom A with its magnetic moment pointing up, state I will be slightly favored over state II. The electronic spin magnetic moment of atom A will have a slight polarization up; that on atom B, down. Therefore a nucleus on atom B will find a nonzero field owing to its own electron. Since this field would reverse if the nucleus on atom A were reversed, an effective nuclear-nuclear coupling results. We can easily estimate the size of the coupling. The fractional excess of state I over state II will be
¥llIeh2JU(O)I~ = hyperflne energy LJE
electrostatic energy
(4.203)
where lu(O)I~ is the wave-function density of the electron lit atom A, and LJE is the energy to an appropriate excited state. The coupling of the electrons on atom B to nucleus 2 are thus given by the proouct of the electron spin coupling if the electron spin is in one orientation only, (81r/3ht1'eIi2Iu(O)11, times the excess fraction of the time the electron is in the favored orientation. Thus the coupling is
(~llle1t2Iu(0)1~) (¥llle1t2Iu(O)11)
(4.204) c,E This coupling turns out to have the correct order of magnitude. If the electron funclions do not contain an s-part, we should instead use the ordinary dipolar coupling between the electron and nuclear spins. The extension of these ideas to solids was made independently by Bloembergen and Rowland [4.19] and by Ruderman and Kittel [4.20]. We shall discuss the situation for metals, confining our attention to the coupling via the s-state hyperfine coupling. Since the metal is not diamagnetic, we should concern ourselves with the possibility of 11 first-order effect - related, therefore, to the Knight shift. This mechanism of coupling was originally proposed by Frohlich and Nabarro [4.21]. As Yosida [4.22] has explained, however, the Frohlich-Nabarro effect is actually included in the second-order calculation. We shall discuss the physical reason shortly, but for the present, we shall simply ignore any first-order effect. Van Vleck [4.23] also discussed this effect. 133
1be effect of the magnetic moment of a nucleus at any lattice site is to make that site a region favorable for an electron of parallel magnetic moment but unfavorable for an electron of anliparallel moment. In order to take advanlage of the magnetic interaction, an electron of parallel moment will distort its wave function to be larger in the vicinity of the nucleus. The distortion is brought about by mixing in other states k of the same spin orientation. As we shall see, the result is as though only states above the Fermi surface were added. TIle wave functions of the Bloch states are added so as to be in phase with the unperturbed function (Fig. 4.13) at the nucleus in order to interfere constnlctively at that point, but because of the spread in wavelengths, they rapidly get out of step as one moves away from the nucleus.
,,-,,
(a)
,
,
~
(b)
"'-----./
x-o
"'-----./
•x
Fig.4.13a,b. The UnllerturbM fun<;lion and two of the higher states mixed in. The nucleull is at z = 0, which gUIlrAntees th"t the WIIYes be in ph"se lit z = O. Nole thl\llhc IIdmixed Waycs beal wilh one Allother. (a) Two of lhe WllVes mixed in by lhe perturbation. (b) Unperturbed function
As a result of the beats between the unperturbed and perturbed functions, the original uniform distribution of spin-up charge density (we neglect the variations due to the lanice charge) is changed to have an oscillatory behavior, which dies out as one goes away from the nucleus. The characteristic length describing the atlenuation is the wavelength of electrons at the Fermi surface. The resulting charge density of electrons whose momenl is parallel to the nucleus is shown in Fig. 4.14.
,
We now tum to the actual calculmion of these effects. For simplicity we shall calculate the interaction between the two nuclei directly rather than compute the changes in the spatial distribution of electron spin. However, the oscillatory natun; of the charge will be apparent from the answer. We consider, therefore, an electron-nuclear coupling 7ten involving only two nuclei of spins II and 12 and for simplicity, treat only the effect of the s-state coupling. We have, then,
7ten = 81\" ['Yl'Yeh211 • 3
L Slb(rl -
R I ) + "I2'Yeh2 12 .
I
L: Blb(r/ -
R2)]
I
= 'Ii, + 'Ii,
(4.205)
where we have allowed the nuclei to be different by using two values of gyromagnetic ratio "II and /2 and spins II and 12. We must lake into account the exclusion principle for the electrons. Two methods are available. We could use (4.205) to find penurbed one-elecrron functions and then fill these functions in accordance with the exclusion principle. Or, we could do penurbation theory in which we used the many electron functions as the unperturbed states. It is this laller procedure that we shall utilize, since it emphasizes that bas.ical1y we are dealing with a many-electron problem and that the states we use are but one approximation. Let us therefore consider a many-electron Slate 10) with energy Eo and excited states In) with energy En. and compute the second-order energy shift due to ?ten. We shall as usual assume the total wave function of the system to be a product of the electron and nuc1e:lr functions. Denoting the latter by 1/;0 with energy Ecn where 0 represents the nuclear spin quantum numbers, we shall wish to compute the second-order energy shift of states 10)"'0'. Therefore, for the second-order shift, we have .6£(2) = 00'
L ",0
.6E~~ of the
(Oal'h'enlna')(na'I'HenIOa) + EO') - (En + Eo-')
slate 10)"'0': (4.206)
,(Eo
Since the electronic energy differences are generally much greater than the differences in nuclear energy, we can neglect Eo and Ecr , in the denominator. 7 Writing 7ten as 'Ht + 7t2, we find
.6E~~ = L
11,0.'
E
I E [(Ool7illua')(ua'I1tJ 100') 0 -
n
+ (Oal7tzlna/)(na'l7izIOa) + (00"11-£ II Ha/)(na'I1-£2100") + (Oo:l7i2Ino:/)(na'[7iIIOa)j (4.207)
o
x
.1g.4.14. The charse delll;ity of eleclrons whose magnetic InOlnClI1$ aN: paral1el to the nuclear moment. The nucleus is located al z o. Po is the charse denllity in the absence or a nuclear moment. At z = z\, the dedron eharse is deficient, lIQ that lhe net electron moment there is oppoosoOO to the nuclear moment
=
134
7 One can think of lhis as being, thel'dore., basically :a calcul:alion of the cOlll'lin.g betW(lCn nuelei in zero external field. However,.t turns olllthat the fidd dependence IS very small. This resull$ from the flld thal the statcs E .. have a continuous distributioll in Cflergy slarting from Ea. Neglecting E.. - t:;. docs not. ~rKnlSly JX'rtur~ lhe.s~tra of excited stales or the malrix elemenl$. A further discuSI510n or the;e f'Omu 15 gIven immediately following (4.223).
135
The first two terms in the brackets represent the changes in energy we should have were onc or the other of the nuclei the only one. The last two terms represent the extra energy when bOlh are simuhaneously present, and are thus an interaction energy. Since it is only the interaction thai we wish 10 calculate, we shall consider the last two terms only. We have then .1E(2) "" ' " (Oal1tdno')(no'j1i21 0o ) +complex conjugate "'"
(4.208)
Eo - £ n
L..n,oJ
Now. fTOm the form of 'HI and 112. we can write them as
L
HI '" I) . Gl;:;
I\fJG11J
L
.jjiif
.Jiif
00'
x
",,(OIGjpln)(nIG,p,IO) ,,( II I ')( 'II I) L E _ J;' L 0' 1/3 0' ex 2{j' 0' +C.C. fJ,P' II 0 LJJj
"" L
"
,,(OIG,pln)(nIG,p' 10) E ~ (alhpI,p'la)+c.c..
/J.P'
n
(4.210)
(nWIO) =
(1,p'
[ " (OIG'P1n)(nIG,p,IO)] E _ J;' + C.c.
11(112(1' L..
..
0
(4.211)
.
LJJJ
(01
1felf =
L 5,6(r, -
L.. fI
Rj )Io)(nl
L 5,6(r, -
, _ E,)
(E 0
I
(-I)('P+P")
=
J
P" P[A'(I)B'(2) . . .j'
VP(A(I)B(2)C(3) ... )dT
-2,
L
(_1)1'"
N. P,P"
= L(-I)I'"
JP"
P[A'(1)11'(2) ... j' V P[A(I)B(2) ... )dT
J
P"(A'(I)B'(2) ... j'V(A(I)B(2) ... jdT
(4.215)
'"'
where the last step follows because V is unchanged by relabelling the electrons. lei us now consider V to be a sum of one-electron operators: V = LV
(4.216)
,
R,)I0)
·h +C.c.
(nWilO) = L(-I)I'"
64.'
C = -9-;11"21';11'· We shall now take the states 10) to be prcxiucts of Bloch functions. Denoting the product of a Bloch function and a spin function by letters A, and so on,
J
P"[.4'(I)B'(2) .. .j' V(1)[A(l)B(2)C(3) . . ·ldT
pit
(4.212)
where
136
(4.214)
where V(f) depends only on the coordinates of the lth particle. Consider, for example, the contribution from l = I:
By expressing the couplings G] and G2 explicitly, we obtain 1 ]. , C"
P'[A'(I)B'(2) .. .j'VP[A(l)B(2)C(3) ... ]dT
,pit
X
" 1felf = L.J
J
-2, L N. P
O.un
To evaluate the energy of (4.210), we should need to specify the nuclear states 10), lust what they would be would depend on the total nuclear Hamiltonian, which would include such things as the coupling of the nuclei to the external static field Ho, the dipolar coupling between nuclei, and so forth. It is convenient to note that whatever the states 10) may be, the energy ~E:;;! is just what we should find as the first-order penurbation contribution of an extra tenn in the nuclear Hamiltonian 1felf given by
(-I)P+I"
Since V is symmetric, it is only the relative ordering of the states that counts. We can express this fact by defining the pennutation pll. pll is the pennutation that, following p, gives the same ordering of electrons as pi alone. That is, pO p = p'. By making this substitution into (4.214), we get
<)',
= L..- L..-
L
;.,.
vN! pp' ,
I,pG,p
where Gj and G2 do not involve the nuclear spin coordinales. By using these
A£(')
P
Of course the pennutation causes any function to vanish if any two functions such as A and B are identical. Consider the matrix element of a perturbation V, which is symmetric among all the electrOns; that is, it is unchanged by interchanging the elecrron numbering:
(4.209)
relations,
(4.213)
jn) = _1_ L(-I)P P[A'(l)B'(2)C'(3) ... )
(J=z,l/,z
.u
p
(nWIO) =
P=Z,II,Z
'Ii, = I, . G, =
10) = _1_ L(-I)P P[A(I)B(2)C(3) ... ]
This will vanish unless the state In) contains B(2)C(3). write such an excited state as
In) = _1_ D-I)P P[A'(I)B(2)G(3) ... 1
.jjiif
.
(4.217)
. Let us therefore
(4.218)
p
This makes 137
(nlVtlO) , I:<-I)P"
J
P"IA'(I)D(2)C(3) ... ]' p" x V(I)IA(I)B(2)C(3) ... ]dT
=0
irA' is idenlicalto any of the funclions othe.,.;,. .
n,e,D.
etc. (4.219)
, (A'IV(I)IA)
It is clear thai the various values of I in (4.216) will simply pick out the various states A. B. C in 10), and the sum over ex:cited stales In) will pick out the states A', n', and so on, which are not occupied in 10). We can therefore write (4.212)
as
'Ii," '
L
C
Now. the Fermi funclions as well as the energies E kll depend on the energy associated with the electron spin quantum number. For example. the electron spin Zeeman energy changes with 3. However, the Fermi levels of the spin-up and spin-down distributions coincide. Thus. at absolute zero. there is a continuous range of Ekll up to the Fenni energy and a cominuous mnge of E k ,.. from the Fermi energy on up. The matrix elements of the 6·functions vary slowly with energy. Therefore the variation of 'Herr with the electron spin energy is very small, and we may forget about the energy of the electron spin coordinate, writing to a good approximation (4.223) in the fonn it would have in zero magnetic field:
'Ii,"' C L
II
k,.s occupied k', s' unoccupied
X
(ksIS6(r - R,)lk's')(k'iIS6(r - R2)ks)
Ek. - £1.:'"
where we have replaced
Eo - En
x
k -
k'
(4.224)
We can now perform the sums over" and s':
by Ek. - £1.:'/," since lhe stales
Eo
if k, s is occupied in State 10)
°
=
·1,
(klo(r - RIl!k')(k'lo(r - R,llk)f(k)[1 - f(e)] E E +C.c.
·12 + c.c. (4.220)
and En differ in energy solely by Ihe rransfer of one electron from the stale lk,,) 10 Ik', S'), The tenns "k, s occupied", "k'. i unoccupied", refer to whether or nOi these siales are occupied by eleclrons in the wave funClion 10). If we define the functions P(k.s) by
p(k, s) = I
II' ('15Ii)(,'151')
kk'; 11,11'
LII·(,151,')(iI5I,)·I" .,11'
L
.,.'
IIp(,15pliXi I5p'I')[w
11"":'""
,L
(4.225)
ltp[,p' T, {5p5p'}
PoP' But, as we saw in Chap. 3,
if k,3 is unoccupied in state 10)
(4.221)
1 , }5(5 + 1)(25 + I)opp'
we can easily remove the restrictions on k,3 from the summation:
0IJIJ'
,--
'Ii,"' C L
since
2
k,lI;k',.'
(4.226)
1 S=-
2
which gives, finally.
P(k,,)(1 - p(k',il] E E ·/z+c.c. (4.222) k. - k'" In order to express the variation of 'Herr with the tempemture of the electrons, we must average 'Herr over an ensemble. This will simply replace IJ(k, 3) by f(k. 3), x
'Heff = I
,C {" (kl<5(r - Rtllk')(k'I<5(r - R,)lk)f(k)(1 - f(e)]
t· .1'2-
2
L. kk'
E k - Ek ,
+ c.c.}
(4.227)
the Fermi function. We have then,
'Herr = C
"
L. k,II;k',If'
= A I 2II'/z
II' (kslS6(r - R J )J'/s')(k's'IS6(r - R:z)lks)
E kll
f(k,,)[1 - f(e,i)] X
-
where AI2 is a constant independent of spin. We now evaluate the matrix elements in terms of the Bloch functions:
E k 'II'
, ·..I2+c.c.
!f;k = 11k ( r ) e ik· r
Eklf - Ek'If'
, C
L
,
where. as before, lIk(r) has the periodicity of the lanice. We have. then,
II' ('15Ii)(iI5I') ·1,
(k'16(1' - R2)lk) = U~,(R2)Uk(R2)ei(k-k')' R]
(kl<5(r - Rtlk')(k'I<5(r - R,)lk)f(k")11 - f(k',i)] X
E
E
kif -
138
+c.c.
k'.'
(4.228)
(4.229)
so that.
(4.223) 139
I: co, I(k -
(klo(r - RI)Ik')(k'lo(r - R,)lk) 2 -R I ) , (R)'(R) (R)i(k-k')'(R = Uk'(R2)Uk' I uk 1 "k 2 e
(4230) .
k,k'
If we assume that R, and R 2 are equivalent sites (as, for example, in a simple mctal), and define
=
(4.231)
k')· RI,]f(k)[1 - f(k')] k2 k,2 -
(-,-)6 2'11'
j'J [COS (kR COS 0) cos (k'R COS 0') k2 _ k,2
+ ~'i:::n-'.(kc:'R::..:c::o,,-';:;0,,):::,':::n.,(;;.k..:'R.:..:c:::o,:.:.O'.!..!)] k2 _ k,2
we have =
HeW
C1112 '. '" 1".,(0)I'I".(0)I'2co, [(k 2
=
E -E,
L-
k,k'
k
64 2 2
I"[r9'Tr 'Yc/1'Y2 h X
k') . R12J f(k)[1 _ f(k')]
E1,;
I.:,k'
Ekt
",
(4.232)
2
j'J sin kR sin k'R kk'f(k)[1 -
4 1 -- (2'11')4 R2
k')· RI,]f(k)[1 - f(k')]
It is not possible 10 evaluate the summation without either some further approximations or some explicit infonnation on the k dependence of wave functions and energy. If one assumes spherical energy surfaces, an effective mass m*, and that !uk/(O)12 and lu/.:(0)1 2 may be replaced by a v;\lue appropri:ltc to k and k' near the Fenni energy, one can evaluate the sums. In these terms,
Ek = 2m*k
(4.236)
f(k)[1 - f(k')jk'k"dfldfl' dkdk'
The integrals over dn = -2'11'd(cos 0) and dn' = -2'11'd(cos ( 1 ) are readily performed to give, for the summution of (4.236),
k
1
I: 1".,(O)l'I".(O)I'co, [(k -
X
k2 _ k,2
f(k')Jdk dk'
This integral may be evaluated at absolute zero by noting that the limits on k' can go from 0 to 00, nO( just kr to 00, since if k' < kp for each k = k J , k' = 1.:2, there is a k = k2, k' = k J which has the opposite sign for the integrand. The range of k l is then extended 10 cover -00 to +00, and the integral is evaluated by a contour integral. The infinity at k = k' is avoided by taking a principal parI. The final result is 2 'Heff = - 9 '11' ,;,1I2 1i2m *!Uk F(0)j4
(4.233)
(sin 2k p R - 2krR cos 2kFR) I
I
x I ' 2
R'
This gives us
'Herr = I, . h fl 2
I: co, [(k I
k,k
k')· R12] f(k)[1 ~ f(k')]
}.;2 _ }.;/2
(4.234)
The number dN of states in k-space within a solid angle ef[} and between two spherical shells of radii k and dk is dn k 2dk dN = 41f 21f 2 Denoting the angle between k and R l 2 as 0, that between Fig.4.15), and using R for IR 121, we have
(4.235)
e and R12 as 0' (see
Fig.4.IS. Relalive orienlations of k, k', and R I 1
140
(4.238)
We note that this expression indeed corresponds to an oscillatory behavior as one varies R. For large distances the coupling goes as
~ 1r 21';'YI1'2 h1 Iu kF(O)]"
X 2m'
(4.237)
(4.239) We see that the dependence on lu(0)1 4 will cause the coupling to be large for large-Z atoms. In fact the coupling for the heavier elements is substantially larger than the direct dipolar coupling. Had we used the conventional dipolar fonn of coupling between the nuclear and electron spins appropriate to "non" s-states, we should have obtained
~ = neff
[r • r _ 3(l1 . R J2)(l2' R I2)] B 1
2
R2
12
(4.240)
12
where B12 is a complicated function. It vanishes if there is no "non" s-character to the wave function as seen by a single nucleus; for large distances, B 12 typically falls off as 1/R12' For a discussion of B 12 the reader is referred to the paper by B/oembergen and Row/and [4.19]. In the case of molecules such as PF3, Gutowsky et al. [4.17] show that the assumption of p-type wave functions on both P and F together with the "non" s-state coupling gives a coupling of the lonn 141
(4.241)
when averaged over all orientations of the molecule in the external field. The coupling does nol vanish when averaged over molecular orientations, since the induced spin magnetization itself depends on the orientation of the molecule with respect to the nuclear spins. Couplings such as (4.240) have the same spin dependence as thai of the direct dipolar coupling. To emphasize this similarity, the coupling is often referred to as the "pseudo-dipolar" coupling. On the other hand, a coupling such as AI2II • h has the same fonn as the electrostatic exchange coupling. Since in our case the physical origin is not an exchange integral, the tenn is referred to as the "pseudo-exchange" coupling. The effect of pseudo-exchange or pseUdo-dipolar coupling on the width and shape of resonance lines can be analyzed by simply adding these tenns to the dipolar tenns. The number of situations that arise are very npmerous. In liquids. the pseudo-dipolar coupling averages 10 zero, but the pseudo-exchange does not, giving rise to resolved splittings. For solids, both tenns have effects. The pseudo-exchange lenn, since it commutes with [~ = [1% + h%, has 110 effect on the second moment of the resonance but does increase the fourth moment. As Van Vleck discusses, this must mean that the central portion of the resonance is narrowed, but the wings are enhanced (see Fig. 4.16), since the fourth moment is more affected by the wings than is the second moment. The fact that the central portion appears sharper gives rise to the tenns exchange narrowing or pseudo-exchange narrowing. Where a real exchange interaction exists. as in electron resonances, the exchange narrowing may be very dramatic.
\
--
Fig.4.16. Solid curve, the resonance shape wilh vanishing pseudo-exchange coupling. The dashed curve shows the shape when pseudo-exchange i" 111· c1uded
analysis of the electron-nuclear interaction. Denoting the wave vector of a Fourier component by q, Yosida points out Ihut the Frohlich-Nabarro effect is the q = 0 leon (infinite wavelength). In the second-order perturbation, a component q '" 0 joins--two electron slates k and k' that satisfy the relation
k' = k + q
Since in a proper second-order calculation the excited and ground states mllst differ, we see from (4.242) that we must exclude q = O. Yosida treats q = 0 by first-order theory, but q f. 0 by second order, adding the results. The answer he obtains in this way is identical to that of Bloembergen-Rowland and Ruderman· Kittel, taking a principal part as mentioned on p. 141. The simplest way to see that the answers will be the same is to consider instead the Knight shifts. There are two ways one can calculate the Knight shift. The first method is to assume a unifonn static field andJO compute the first-order nuclear-electron coupling of the polarized electron state. The second method is more complicated. We assume a static field but one that is oscillating spatially with wave vector q. For such a spatial oscillatory field there is no net spin polarization, since the field points up in some regions of space as much as it points down in others. The usual first-order interaction vanishes. If one now goes to second-order perturbation in which one matrix element is the electron-nuclear coupling, the other, the electron-applied field interaction, a nonzero result is found. That is. the static field induces a spatially varying spin polarization. Let us choose the maximum of the static field 10 be at the position of the nucleus. When q gets very small (long wavelength), we expect that the result must be the same as if the field were strictly unifonn. Therefore the limit of the second-order answer as q ...... 0 must be the usual Knight shift. This result can in fact be verified. In evaluating (4.238) using principal parts. one takes the limit as k ' --+ k. This procedure, as with the Knight shift, includes the first-order perturbation repopulation contribution. or the Frohlich-Nabarro effect. Before concluding this chapter, we should consider the role of electron spin in the chemical shift of diamagnetic substances. In the absence of an applied magnetic field, diamagnetic substances are characterized by 11 total electron-spin quantum number S of zero. The application of a field Ho in, say, the z·dircction adds a tenn Hsz. the spin Zeeman interaction, to the Hamiltonian:
If the two nuclei are not identical, one can approximate the pseudo-exchange ~o~pling as AI2[lz[2z' Since this interaction does not commute with 11%
+ h%.
11 ltlcreases the second moment. The resonance curve then appears broadened, and one speaks of "exchange broadening". If there is a quadrupole interaction that makes the various m-states unequally spaced, the exchange coupling can lead to a broadening. even when the nuclei are identical. We conclude by remarking on the Frohlich-Nabarro effect that the presence of one nucleus causes the electrons to repopulate their spin states, producing a magnetic field at other nuclei in much the same way as the static field produces a Knight shift. Yosidtl has analyzed the problem by perfonning a spatial Fourier 142
(4.242)
Hsz
= 'YehHo
E" $zj = 'YehHoSz
(4.243)
j=1
where j labels the electrons, and where N
S,"
L: S,;
(4.244)
j=1
Since the ground-state wave function 10) is a spin zero function. we have that
S,IO) =
°
(4.245) 143
so that all matrix elements of 'H.sz to excited states In) vanish:
°.
(4.246) (nl1iszIO) = 7.hHo(nIS, 10) = The ground state is therefore strictly decoupled from all other states as far as the spin Zeeman coupling is concerned. 1be applied field is therefore unable to induce any net spin, and there is no phenomenon analogous to the unquenching of the orbital angular momentum. The fact that the spins actually couple to a magnetic field makes this result seem strange. Intuitively we except that, given a strong cnough magnetic field, the spins must be polarized. The parndox is resolved by considering an example, the hydrogen molecule. The ground state is the singlet bonding state, but there is a triplet antibonding state. In the presence of an applied field, the states split, as shown in Fig. 4. 17. Energy
Triplet
Singlet
S-l,M._O
S.O,M.-O
Ag. 4.17. Erred of lhe applied field /10 on the singld. And triplet-spin stales of a hydrogen mole<:ule. If 110 were large enough, a triplet slate would be lowesl Ilnd the ground state would pOSSCliS & magnetic moment
H,
As we can see, for large enough Ho the S = I, /tI/s = -I state crosses the S = 0, Ms = 0 state. The ground state is then a triplet state, corresponding to a spin polarization. However, since the singlet-triplet splitting in zero field is several electron volts, the crossing of levels could never be produced by an auainable laboratory field. Further insight is obtained by considering the effect of a hypothetical mixing of a triplet state into the ground state. If the result is to induce a net spin polarization in the positive z-direction on one atom, it induces an equal and opposite spin polarization in the negmive z-direction on the other atom. Clearly there is zero net spin Zeeman interaction with an applied field in the z-direction. Since such a spin polarization gives no net lowering of energy, it is not in fact induced. Note, however, that if the two atoms are dissimilar, there may be a different induced orbital moment which, through the spin-orbit coupling, could then induce such a spin polarization. Thus, in a molecule such as H I, the iodine orbital magnetization could induce a spin polarization into the bond, giving a spin contribution to the chemical shift on the hydrogen as well as on Ihe iodine. 144
5. Spin-Lattice Relaxation and Motional Narrowing of Resonance Lines
5.1 Introduction We turn now to a discussion of how the nuclei arrive at their thermal equilibrium magnetization via the process of spin-lanice relaxation. We shall find it convenient to cliscuss two techniques for computing TI· The first method is appropriate when the coupling of the nuclei with one another is much stronger than with the lanice. In this case an allempt to compute the population changes of an individual nucleus due to th~ coupling to th~ lattice is complicated by the presence of a much stronger coupling of the nuclei among themselves. The first method makes the assumptio.n that the strong c~u. piing simply establishes a common temperature for t.he SpillS and that the .laHlce coupling causes this temperature to change. There IS a close analogy wllh the process of heat transfer between a gas and the walls of its contain~~, i~ which the role of the collisions within the gas is to maintain a thermal eqmhbnum among the gas molecules. In the collision of molecules with the wall, we con~i~e~ the molecules to have the velocity distribution appropriatc to themlal eqUlhbnum. As we shall see the first method leads to a fomlUla for TI that is particularly convenient when the lattice is readily described in quantum mechanical tenus. For example, relaxation in a metal involves the transfer of energy to.the conduction electrons, which are readily thought of in terms of Bloch funcuons and the . ' exclusion principle. The second method is that of the so-ealled density matnx. Although thiS is a completely general method, it finds its greatest utility for systems in w~ich the lallice is nalurally described classically and in which the resonance WIdth is substantially narrowed by the motion of the nuclei. Moreover, when motio~ takes place, the relaxalion time T2, which describes coupling between the nUcl~I, becomes long, and it may be a very poor approximation to assume that a SplO temperature is achieved rapidly as compared to T l . lllUS the second ~ethOO is useful when the first one fails. Because of their large mass, the motion of the nuclei is often given very well in classical tenns. In fact an attempt to describe the motion of molecules in a liquid quantum mechllOically would be quite cumbersome. Consequently the density matrix. method is well suited t.o discussing cases in which motional narrowing takes place. An added feature IS that both TI and T2 processes (relaxation of I z and I z or III) can be treated by lhe density matrix method, provided there is motional narrowing. 145
The density matrix method is very closely related to the conventional timedependent perturbation theory. Actually the two are entirely equivalent. The density matrix method, however, gives results in a particularly useful form. It is ideal for treating problems in which phase coherence is important, and in fact the density matrix or a mathematical equivalent is necessary to treat such problems. In any event, much of the fonnalism of this chapter applies equally well to systems other than spins. For example, dielectric relaxation can be treated by these methods. As we see, the two approaches complement one another, one applying to the broad resonances of a rigid lattice and the other being most useful when the resonance has been narrowed by nuclear motion.
would enter only when we tried to express the wave function of the total system in terms of the wave functions of individual spins. We shall say that any system whose population obeys (5.1) is described by a temperature T, even when the system is nOt in equilibrium with a reservoir. Equation (5.3) enables us to make a simple plot to schematize the populations. We illustrate in Fig.5.1 a case based on our first interpretation: we consider the population of the various energy states of a single spin of I = acted on by a static magnetic field.
t
,
I
I
5.2 Relaxation of a System Described by a Spin Temperature I
P(Ea )
e-E~/kT
P(EI1) - e E.lkT
so that, since
(5.1)
E.
we have P(Ea )::=.
e-E.. IJcT
L: e ,
Ec/kT =
e- E.. IJcT Z where
(a)
(b)
.'ig. 5.1. (a) Energy levels ar II spin 3/2 nucleus. (b) Bllr grlll>h af POpullllion vel'$US energy.
The lengths or the bal'S are determined by the exponential em'e1ope
Energy
(5.2)
LP(E.) = I
,
+% +% - - -
I
Exponential I envelopo ".i=-~
-'h
A system with a set of energies E u , Ell' and so on, which is in thermal equilibrium with a reservoir of temperature T, occupies the levels with probabilities p(Ea ), p(E/I), and so forth, which arc given by
I
I
- 'It -1ft, (5.3)
'I, 'I,
Population
Ca)
Cb)
(5.4)
is the partition function or "sum of states". These equations may actually have two interpretations, which we might illustrate by considering N identical spins. The first interpretation considers the spins as isolated from one another. The "system" consists then of a single spin, and the energies E a represent the possible energies of the single spin. The second interpretation considers the system to be formed by all N spins. In this case, E a represents the total energy of all N spins. We shall find it convenient to use both interpretations. We should perhaps note that the first interpretation is correct only if the spins can be considered to obey Maxwell-Boltzmann statistics, . but the second interpretation holds true whether or not the individual particles obey Maxwell-Boltzmann, Fermi-Dirac, or Bose-Einstein statistics. The statistics
1<1g. 5.2. (a) Population distribution not describable by a temperature. (b) A po56ible lransition or a pair or spins rrom the st...tes designaled by CI"OMes to lhose designated by circles
A system such as shown in Fig. 5.2 clearly does not correspond to thennal equilibrium, since the bar graph envelope is not an exponential. In Fig.5.2b we indicate a transition that could take place, conserving the total energy of the spins. Two spins designated by crosses couple, inducing transitions to the states designated by circles, one spin going up in energy and the other down. (Such a transition would be induced by the Ii tenns of the dipolar coupling.) The number of transitions per second from cross to circle, dNldt)r ..... O. will be the product of the probabilities of finding two spins in the initial state times Ihe probability of transition W., ..... 0 if the spins are in the initial state. Thus,
It
I See rererences to "Spill TemperMure- in the Bibliography
146
147
~)
._0
= P_I/2P_I/2W,;_O
(5.5)
The inverse reaction from the circle 10 the cross will have a rale dNldOo .... ~, given by
dN)
dt
0-.
=:
(5.6)
P_J/2PI!2 WO_J:
If we equate these rates, we guarantee equilibrium. This is the assumption thai equilibrium is obtained by detailed balance. Since WO_:t: "" W:>: ..... o, we find P-I/2P-I/2 = P-3/2PI/2
P-3/2
P-I/2
P-I/2
1'+1/2
or - - = - -
(5.8) n
The average energy of the system, E, is then (5.9)
n
We shall assume further that the energies E u nre measured from a reference such that
L;En =Tr1i=O
(5.10)
,
n
a condition that is fulfilled for both the Zeeman and dipolar energies. To compute the relaxation, we shall consider changes in the average energy. If we define fJ = l/kT to represent the spin temperature, we have that
dE dE dP di"=dpdi"
dpni ::::: "" d LJ(Pm W mn - pu W nm )
(5.11)
'"
dE dt
m,n 1 = '2 L(PmWmn - PnWnm)(En - Em)
(5.14)
'",n where the second form is introduced because it treats the labels m and 11 more symmetrically. By equating (5.11) and (5.14), we obtain a diITerenlial equation for the changes in spin temperature. We have two problems: (I) finding dE/dfJ and (2) seeing what becomes of (5.14) when we introduce the requiremenl that at all times a spin temperature apply. We tum first to evaluating dE/dP: e-fjE.
Pn = - -
Z
dE
and
(5.15)
d
dP = dP ~Pn(P)En
(5.16)
We first seek an approximate expression for Z. Once again, assuming the temperature to be high enough so that PEn <:: I for the majority of states, we expand exp( -PEn) in a power series and keep only the leading terms:
P2E~)
Z= ~ ( l-fJEn+--:v:- + ...
(5.17)
Approximating such a power series expansion by the leading terms clearly has validity if IEnl <:: kT for the significant energies. However, the approximation proves legitimate under less stringent conditions using an argument similar to that in Appendix E both here and in (5.19-26). When we utilize (5.10), the second lenn on the right of (5.17) vanishes. We then neglect the f32 term, and Z becomes equal to the total number of states. Since this is also Z for infinite temperature Zoo, we may say that
Z = Zoo
But since E::::: EnPnEn, we have also thai
(5.13)
This equation is frequently called the "master" equation. By substituting into (5.12) we have, then, that
(5.7)
But this is just the condition of thermal equilibrium among the states, since they are equally spaced in energy. We see, therefore. thai thennal equilibrium is reached by processes such as we have indicated in Fig.5.2b. The typical rate for such a process is of the order of the inverse of the rigid lattice line breadth, or between 10 10 lOO/tS for typical nuclei. Therefore, if Tl is milliseconds to seconds, we should consider the nuclear populations 10 be given by a Boltzmann distribution. We shall now proceed to consider the relaxalion of a system of nuclear spins whose Hamiltonian 1t has eigenvalues En, and in which the fractional occupation of state n is Po. (Thus n designates a state of the total system, rather thall the energy of a single spill). Nonnalization requires thal
E=LPrIEn
We shall assume that the Po's obey simple linear rate equations. Introducing Wmn as the probability per second that the lattice induces a transition of the system from m to n if the system is in state m, the rate equation is
(5.18)
When we utilize this fact and (5.15), (5.16) becomes
de : : : dt '48
d "" dPn dt LJPoEo ::::: "" LJEndi n
(5.12)
n
149
where
L =
Wmn(E m - E n )2
1 m,1I
2
(5.28)
L:E~ n
-;::t
(5.19)
I " 2 -ZLJE" n
00
again in the high-temperature limit. Thus
" E' dj3~n
dE
(5.20)
dt<::-dt~
We now tum to evaluation of (5.14). Since the system is always describable by a temperature, we have PII
_ P e(Em~En){J
(5.21)
m
-
We shall furthennore assume that when the system is in thennal equilibrium with the lattice, the transitions between every pair of levels are in equilibrium. This is the so-called principle of detailed balance. Denoting by p~ the value of Pn when the spins are in thennal equilibrium with the lattice. the principle of detailed balance says thai
LWmn = PIIlOWnm Pm I.
P w:mn-- W nfllL ---!:!... Pm
(5.22)
or that
5.3 Relaxation of Nuclei in a Metal
= W nm e(Em-En).B"
(5.23)
where PI. = I/(kTL)' By subsliluting (5.21) and (5.23) into (5.14), we find
dE dt
=.!. L: pm W mn [1
- e(Em-En)({J-fJ!.l](E" - Em}
(5.24)
2 m,n
Now, expanding the exponential, we find dE dt
(5.25)
Now pm=
e- f3Em Zoo
~
1 - (3E m + (/32 E~/2!) ....., =
Zoo
Zoo
(5.26)
Thus, combining (5.26) with (5.25) and equating the resultant dEldt to that of (5.20), we find
dP dt
L: Wmn(Em -
=
En)']
(P _ P) ~ ,"m""n_~::;;;- _ _ L
[2
Equation (5.28) was first derived by Gorter on the assumption that 1(3 - (31, 1« (3 [5.1]. As we can see, this restriction is not necessary. The great advantage of (5.28) is that, by postulating a temperature, it has taken into account the spin-spin couplings. The rate equations, (5.14), would by themselves imply that there are multiple time constanlS that describe the spinlattice relaxation, but the assumption of a temperature forces the whole system to relax with a single exponential. We may get added insight by viewing our states n as being nearly exact solutions of the nuclear spin Hamiltonian, between which (since they are not exact states) transitions take place rapidly to guarantee a thermal equilibrium, but between which the lanice also makes much slower transitions. After each lattice transition, which disturbs the nuclear distriblllion, the nuclei readjust among their approximate levels so that the lattice once again finds the spins distribllled according to a temperature when it induces the next spin transition. Our formalism implies that treating the states n as being exact makes a negligible difference in the answer.
LE~
(5.27)
We now turn to an example of the application of (5.28). We shall consider the relaxation of nuclei in a metal by their coupling to the spin magnetic moments of the conduction electrons. This is the dominant relaxation mechanism. In a T, process, the nucleus undergoes a transition in which it either absorbs or gives up energy. In order to conserve energy, the lanice must undergo a compensating change. For coupling to the conduction electrons, we may think of the nuclear transition as involving a simultaneous electron transition from some state of wave vector k and spin orientation .5, to a state k l , l. We may think of this as a scattering problem. Denoting the initial and final nuclear quanlum numbers as m and n, respectively, we have that the number of cransitions per second from the initial state of nucleus and electron Imks) to the final state Ink'i), W mkB,lIk' B" is W mkB,nk'B' =
2;
2
l(mkslVlnk' i)1 o(Em + E kB - En - Ek'B')
(5.29)
where V is the interaction that provides the scattering, and where (5.29) assumes that there is an electron in Iks) and there is /lone in Ikll). The total probability per second of nuclear transitions is obtained by adding up the IVmks nk'B' 's for all initial and final electron states. We have '
n
150
151
L
W mn '"
(5.30)
Wmks,Ilk's'
x
ksoccupicd 1,,'6' unoccupied
L
(5.31)
vVmb ,llk's'Pks(l- Pk'6')
ks;k's'
By averaging (5.31) over an ensemble of electron systems, we simply replace Pb by the Fermi function f(Eb), which we abbreviate as f(k,s):
L
W"'>I '"
(5.38)
and, assuming E ks '" E", + E s • we get for W mn
We must now express W mb nl/6' explicitly. To do so, we must specify the interaction V. For metals with a substantial s-character to the wave function at the Fenni surface, the dominant contribution to V comes from the s-state coupling between the nuclear and electron spins:
8. , V'" "3,e,nh 1·50(1')
Imks) '" Im)ls)uk(r)e ik • T
(5.34)
It is a simple matter to compute the matrix element of (5.29):
(mkslVlnk's') '" 81r Iclnh2(mII[n). (sISls')uk(O)uk'(O)
(5.35)
3
which gives us
2:
(mII.ln)(nllo,lm)
(5.36) x o(E", + Eb - En - E k '6') We can substitute this expression into (5.32) CO compute W mn . We are once again faced with a summation over k and k ' of a slowly varying function. As before, we replace the summation by an integral, using the density of stales g(Ek , A) introduced in Sect. 4.7. This gives us
r; -9-'C Inti
'"'
L.
0I,OI'6,S'
(mIIOI[n)(nllo,[m)(sISOI[s')(s'[So'ls)
01,01'6,6'
x
J<1"k(O)I')
X
e(Ek)e(Ek ,)dEk
E
J:
0"",(0)1') E f(E b '"
Il
)1I -
f(Ek", + Em - E,,)] (5.39)
~,~~+~-~,+~-~
(5~
Since E", - En. the nuclear energy change, is very small compared with kT, the Fenni function f(Ek6 + Em - Ell) may be replaced by f(Eb)' Actually, doing so makes W mn = W nm . It is, in fact, at this point that the slight difference between W mn and Wmn arises. It is the slight difference, we recall, that gives rise to the establishment of the thennal equilibrium nuclear population. We are allowed to neglect the difference here in computing W"'>I' since we have already included the effect in (5.23). Moreover, since both u(Ek ,) and ([uk,(0)1 2 ) Ell are slowly varying functions of E k" we may set them equal to their values when Ek '" Ek" In fact we shall evaluate U(Ek ,), and so on at E ks ' This gives us for the integral in (5.39):
J
x (,15.1,')(/150,1,)I"k(O)I'I".,(O)I'
2 2 4
6~2 1;,~h'l L
=
OI,OI'=Z,Y,z
27l' 647l'2
Wmn '" 2;
(5.33)
where we have chosen the nucleus I to be at the onglO. For the electron wave function, we shall take a product of a spin function and a Bloch function uk(r)exp(ik· r). Therefore the initial wave function is
I
I
(mIIOIln)(nIIOI,[m)(s[SOIls )(s 1501 , Is)
f(E)JdE
(5.41)
o where we have set the lower limits as zero, since the only contributions to the integral come from the region near the Fenni surface. E'" E F . Since (5.41) is independent of the spin quantum numbers sand S', we may now evaluate the spin sum of (5.39):
2:('15.1,')(/150,1,) = 2:(,150 5.,1,) S,6'
6 = Tr {SOISOI' } = 00101' !S(S + 1)(2S + I)
'" 00101,/2 152
(5.37)
We integrate first over dA and dA ' , utilizing the relations of (4.153, 154) to perform the integrals, and introducing again the average of [uk(0)1 2 over the energy surface, Ek' (IUk(0)[2) EJ:' We also assume that the energy Eks appearing in the Fenni functions remains constant on a surface of constant Ek' an assumption that would be fulfilled unless the spin energy depended on the location on the surface Ek . The integralion over dEI,,' is then easy because of the delta function. We have that
(5.32)
Wmks,nk's,f(k,s)[l - f(k',s/)]
k6;k's'
W"'>I '"
f(k, ')11 - f(k', ,')Jg(Ek • A)
x g(Ek " A')o(Em - En + Eks - Ek'6,)dEJ.,dAdEk'dA'.
The sum over "ks occupied" is, of course, equivalent to summing over electrons. We can remove the restrictions on ks and k's' by introducing the quantity Pks' which is defined to be unity if ks is occupied; zero, otherwise. This gives us W mu '"
Jl"k(O)I'I"k,(O)I'
since
5 '"
!
(5.42) 153
This gives us
W mn
::::
W mn ::::
~ 1l'3h3'Y~'Y~ L(mIIQC[n)(n[Iolm)
•
x
J(IUk(O)I')~.'(E)f(E)[1
- f(E)JdE
(5.43)
Now f(E)[1 - f(E)} = -kT :~
(5.44)
which follows directly from the fact that
feE) =
~4 1l'31l3'Y;'Y~([uJ.:(0)12)h,£l(EF)kT L
•
L ](mIIQC[n)1 2
W mn =
1
•
L.;j L(mll;.ln)(nllj.lm) iJ
Fig. 5.3. f'mdions j(E), t -j(E), and j(E){I - j(E)I. The $Olid line .how. I(E)(I -/(E)I
Since /(0) = 1 and /(00) :::: 0, and since /(£)[1 - /(E)] peaks up only within a width kT, it is also closely related 10 a 5-function when in an inlegral of other functions that vary slowly over kT.
(see
where the coefficient i = j is aoo, and where aij for i 1: j falls off rapidly with the distance apart of the nuclei i and j. 11lese terms arise because the elecrron wave function extends over many nuclei, so that more than one nucleus may scatter the electron from a given initial 10 a given final state. By returning to (5.47) and employing our fonnula for T I , we have
I TI =
SF
L
_
",pC
I
J(f.,-E~,)j(1-J)dE+dE7
dG 00 , +dE SF 0
154
(ml[?t.I.J1n)(nl[?t. l.Jlm) __=-=::_
a~ cm~,~n~,.,---
+ Ep
00
.
j(E-Ep )2jCt-ndE ...
S 0
EE;n 00
J G(E)j(E)[1 -j(E)]dE = G(E~') J j(E)/1 - jCE)ldE
I
EE;n
(5.45)
7 Equation (5:4.5) can be derived simply. Let C(E) be a slowly varying function of energy. Then, utlh1.lIIg Lhe fact that j(E)[l-j(E)J is nonvanishing within only kT about Ep, we c~n expand C(E) in a power series "bolll EF:
Thus 00
En)'
00
= _
2 2 +(E-Ed d CI 2! dE2
-
~=------
aoo Z-
aoo
"
(mll.lnXnll.lmXEm
'00 •• ~,n~,~.
By utilizing this fact, we have, finally,
G(E)=GCEFl+CE-EddC[ dE
(5.48)
0
L
footnole 7 ).
(5.47)
,
defining the quantity aoo, which is independent of the nuclear states nand m. If we have more than one nucleus, it can be shown [5.2] that Hlmn is given by a sum over the N nuclei, labeled by j or j, (i,i:::: 1 to N):
and f(E)[1 - f(E») peaks up very strongly (Fig. 5.3) when E " E•.
f(E)[1 - feE)] = kT6(E - E.)
(5.46)
We note that Wmn is proportional to the temperature T. This fact has a simple-physical interpretation. When the nuclei undergo a transition, they give an energy to the electrons that is very small compared with kT. Most of the electrons are unable to take pan in the relaxation because they have no empty states nearby in energy into which they can make a transition. 11 is only those electrons in the tail of the distribution that are important. Their number is proportional to kT. We can write (5.46) as W mn :::: aDO
I e(E-Ep)lkT + I
[(m[Io [n)[2
=-2
o=::e,y,::
Tr {?t'}
(5.49)
A similar expression is found by using (5.48). The important point is that we do not need to solve for the explicit eigenstales and eigenvalues, but only evaluale the traces in a convenient represent:ltion.
The first term, using C5.44), gives G(EF )kT. The second term vanishes, since the integrand is an odd function of E - EF, and the third term gives a contribution proportional to (kT)3, as seen by changing the integrand from E lo ::e := ElkT. If we neglect the third and highcr terms, the answer is just what we should have if we replaced j(1 - n by kT6(E - Ep). The correction!! are generally of order (kTIEd 7 smaller, the exact form depending on the functional dependence of C on the energy E.
155
For our problem of a single spin, the quantum numbers m and label the 2I + 1 eigenstates of I r • Then, using the fact that
11
would (5.59) (5.50)
[Ir , I z ]
:::
iIy
(5.51)
etc.
we find
L Tr {['H, IO']2} ::: _1'~h2 H6 Tr {(I; + I;)} • E Tr {'H 2 } ::: 1~'12 H6 Tr {I;} •
(5.52)
Thbl~
so that, since Tr{I;} ::: Tr {I;} :::Tr{I;J,
L a
Tr {[?i.E.I'
:::-2
Tr (11')
(5.53)
~:::: aoo:::: 64 'lf3h3;:;~(luI(O)I)}Fe2(Ef)kT 9
TI
(5.54)
The quantity (lul;:(O)l)}; appearing in the expression also occurred in the expression for the Knight shift, LJ.HIH :
LJ.H
H·
h
2
S
T(lu.(O)I> E,X.
.
(5.55)
We can therefore use (5.55) to evaluate (Iuk(O)!fh., giving
TI
(LJ.:Y : : [£>{~F)r 1f~T 1'~~ih3
The T appearing in the Korringa relation represents only one contribution to the relaxation time - that due to the coupling of nuclei to the magnetic moment of ,!I-state electrons. One expects that the experimental TI should, if anything, be shoner. It is therefore interesting to examine a table given by Pines {SA]. In it he lists e;Jlperimental Tj's, those computed from (5.58) (the Korringa relation) and those computed from (5.59), using Pines' theoretical values of relx& and I!O(E.)/e(E.).
(5.56)
5.1. Experimental and Lheoretical T I '. (all tillW$ in ms)
T 1 (Experimental)
T. (Korrillga) T. (Pines)
t5O±5 15.9±0.3 2.;5±0.2 3.0±0.6 6.3±0.1
88 10.3 '.1 '.3 '.1
232
18.1 2.94 4.'
•••
We nOle that the Korringa TI'S are all sllOrur than the e;Jlperimental ones. The discrepancy cannot be removed by appealing to other relaxation processes, since if we included them, the theoretical T, would be even shorter than those computed by the Korringa relation, and the discrepancy would be still greater. On the other hand, the Pines' values, based on inclusion of the electron-electron couplings, make the predicted values longer than the experimental. The discrepancy between the Pines' values and the experimental is perhaps a measure of the imponance of relaxation processes we have not computed.
For a Fenni gas of noninteracting spins, one can show that X~ given by
5.4 Density Matrix - General Equations
2h2
1 Xo:::: Teo(Er)
(5.57)
where we have put subscripts "0" on XS and e(E F ) 10 label them as appropriate to noninteracting electrons. In this approximation one has
C>H)2
TI ( - H
h
"(2
• - - --"
41fkT "(~
(5.58)
Equation (5.58) is commonly called the "Korringa relation", after Dr. J. Korringa who first published it [5.3]. It provides a very convenient way to use measured Knight shifts to predict spin-lattice relaxation times. A more accurate expression is obtained from (5.56) and (5.57):
156
As we have remarked, the concept of a spin temperature is nOl always valid. We lum now to discussion of a mel hod of attack that is very useful when the spin temperature concept breaks down - the technique of the density matrix. An exceptionally good discussion of lhe densily matrix is that of Tolman [5.5}. The method has the further advantage of giving one a discussion of both T l and T2 processes in a natural way. It is ideally suited 10 treating problems in which the resonance is narrowed by lhe bodily motion of the nuclei. It is also applicable to broad line spectra, whcre it can in fact be used for an alternate derivation of our equation for TI of Scct.5.3. As we shall see, the method is simply a variant of the usual time-dependent perturbation theory, but one that is in a panicularly useful fonn. 157
We begin by considering a system described by a wave function t/J, at some instant of time. and ask for the expectation value (M~) of some operator such as the x-component of magnetization. M~. We have. then, ~
(M.)
Therefore
PMztln =
L PUm(mIM~ln) m
(5.60)
(,p, M.,p)
SupJXlse we now expand in a complete set of orthononnal functions Un, which are independent of time:
(5.69) m,n' so that (5.70)
(5.61) If
t/J
"
m
By using (5.62), we have that
varies in time, so must the cn's. In tenns of the functions
tin,
we have
(M.) ~ I:(nlPlm)(mIM.ln)
(5.62)
m,"
lI,m
If we change the wave function, (M~) will differ because the coefficients C~.CII will differ, but the matrix elements (mlMz In) will remain the same. Correspondingly, for a given ¢, the effect of calculating expectation values of different operators is found in the different matrix elements, but the coefficients c;,lc" remain the same. We can conveniently arrange the coefficients cnc;" to fonn a matrix. We note that, to compute any observable, we can specify either all the c" 's or all the products c"c:n. However, since we always wish the c's in the fonn of products. to calculate observable properties of the system, we find knowledge of the products more useful than knowledge of the individual c's. It is convenient to think of the matrix cncin as being the representation of an operator P, the operator being defined by its matrix elements: (5.63)
(nIPlm) = Cnc:n In tenns of (5.63) we have. then,
(M.) ~ I:(nlPlm)(mIM.ln)
(5.64)
lI,m
The result of the operator P acting on a function
tim
may be written as (5.65)
~DnIPM.ln)
"
(5.71)
Also we note that P is an Hennitian operator. This we prove by noting that the definition of an Hennitian operator P is
JU~PUmdT ==
J(Pun)*UmdT =
(J u~IPu
.. dT
r
(5.72)
0'
(nIPlm)
~
(mIPln)'
(5.73)
But (nIPlm) = CIIC:"
(mIPln) = cmc~
(5.74)
so that (5.73) is satisfied. Often we shall be concerned with problems in which we wish to compute the average expectation value of an ensemble of systems. The matrix elements c"c:n will then vary from system to system to the extent that they have differing wave functions. but the matrix elemems (mIM:r:]n) will be the same. If we use a bar to denote an ensemble average, we have, then,
(Mz ) =
L Cnc~l(mIMz]n)
(5.75)
",m
since the un's fonn a complete set. As usual, we find the an's by multiplying both sides from the left by Uil and imegrating:
an
J
= U~P1LmdT = (nIPlm)
so that
(5.66)
The quantities C"C;" fonn a matrix. and it is this matrix that we call the "density matrix". We shall consider it to be the matrix of an operator e, defined by the equation
(n]elm) = cnei'll - (nIPlm) (5.67) n
Likewise we have M~ltll = LUm(mIM~ln)
(5.68)
(5.76)
Since P is an Hennitian operator, it is clear that e is as well. Equation (5.64) becomes, then,
(M.) ~ I:(nlelm)(mIM.ln)~T'{eM.}~Tc{M.e}
(5.77)
lI,m
m
158
159
For the future, we shall omit the bar indicating an ensemble average to simplify the notation, but of course we realize lhat whenever the symbol (J is used, an ensemble average is intended. Of course the wave function t/J, describing whatever system we are considering, will develop in time. Since the 11 11 's are independent of time, the coefficients C n mUSI carry the lime dependence. It is straightforward to find the differential equation they obey in terms of Hamiltonian 1i of the system, since
_'!. a~ , at
=
'H~
(5.18)
The density matrix is the quantum mechanical equivalent of the classical density (J of points in phase space, and (5.82) is the quantum mechanical fonn of Liouville's theorem describing the time rate of change of density at a fixed point in phase space. In the event that 1i is independent of time, we may obtain a formal solution of (5.82), (](t):: e-(i/l)'ltI{](O)e(i!l)UI
In terms of functions have, for example,
J :: J
which gives, using (5.61),
11
(k1{](t)lm) ::
dCn
--;-1 Ln -dUn:: Lcn'Hu .. t .. We can pick out the equation for one panicular coefficient, both sides by uk and integrating: h dCk -7 -
I
dt
=
2:Cn(kl1tln)
CI;,
by multiplying
U'I>
(5.83)
which are eigenjimcrions of Ihe Hamiltonian 'H, we
uke-(i/')'H'{](O)e(i/l)'ltfumdr (e(i/l)'HI Uk ) •{](O)eWA)'HIUmdr
(5.84)
By utilizing the fact that HUm:: EmU"" and using the power series ex.pansion of the exponential operator, we get
(5.19)
(5.85)
n
This equation is the well-known starting point for time-dependent perturbation theory. We can use (5.79) to find a differential equation for the matrix elements of the operator P, since
."
= -,i
E [ckc;.(nIHlm) -
(5.80)
(kl'Hln)cnc:"l
for the time-dependent matrix element in lerms of the matrix element of t! at t :: O.
So far we have talked about the density matrix without ever exhibiting explicitly an operator for e. For the sake of concreteness, we shall do so now. We shall take an example of a spin system in thermal equilibrium at a temperalUre T. We shall take as our basis states, Uri, the eigenstates of lhe Hamiltonian of the problem, 'Ho. The populations of the eigenstates are then given by the Boltzmann factors, giving for the diagonal elements of (!:
e- E ... / kT
CmC~ :: =--'0--
Z
i
(5.86)
where, as usual,
= h(kl P1t -1tPlm)
Z = 'Le-E./kT
n
where we have used (5.70) for the last step. We can write (5.80) in operator form
as
dP
ill =
i
hlP, 1t1
If we wrile
(5.81)
This equation looks very similar to that of (2.31) for the time derivative of an observable, except for the sign change. If we perfonn an ensemble average of the various steps of (5.80), assuming 'H to be identical for all members of the ensemble, we find a differential equation for the density matrix. (J. Since the averaging simply replaces P by {!, the equation for (J is (5.82)
160
Cn ::
lcllieia-n
we have that CmC~:: ICmIICnlci(Q'... an)
(5.87)
It is customary in statistical mechanics to assume that the phases all arc statistically independent of the amplitudes Ie" I and that, moreover, 0'", or On have all values with equal probability. This hypothesis, called the "hypothesis of random phases", causes all the off-diagonal elements of (5.87) to vanish. If, for example, we were to compute the average magnetization perpendicular to the static field 161
for a group of noninteracting spins, as we did in (2.88), the vanishing of the off-diagonal elements of !! would make the transverse components of magnetization vanish, as they must, of course, for a system to be in thennal equilibrium. More generally, we see from (5.85) Ihal the off-diagon..l elements of fI oscillate harmonically in time. If they do not vanish, we expect that there will be some observable property of the system which will oscillate in time according to (5.75). BUI we should then not have a true thennal equilibrium, since for thennal equilibrium we mean that all properties are independent of time. Therefore we must assume that all the off-diagonal elements vanish. Note. however, from (5.85) (which applies to the situation in which the basis functions are eigenfunctions of the Hamiltonian) that if the off-diagonal elements vanish at anyone lime, they vanish for all time. We have, therefore, (nlelm) = (omll/ Z )e-En /k'l"
(5.88)
It is worth noting that the operator for f! is on a different fOOling from most other operators such as that for momentum. In the absence of a magnetic field, the latter is always li.\J/i. For a given representation the density matrix may, however, be specified quite arbitrarily, subject only to the conditions that it be Hennitian, thai its diagonal elements be greater than or equal to zero. and lhat they sum to unity. There is therefore no operator known a priori. However, in certain instances the matrix elements (nJelm) can be obtained very simply from a specific operator for fl. When this is possible, we can use operator methods to calculate properties of the system. We now ask what operator will give the matrix elements of (5.88) bearing in mind that the u,,'s, and so 011, are eigenfunctions of 'HO. Using the fact that
e
-"Ho/kT
um=e
-Em/kT
Urn
(5.89)
(which can be proved from the expansion of the exponentials), we can see readily that the explicit fonn of e is
e = ~e-"Ho/kT
(5.90)
Z
We can use this expression now to compute the average value of any physical property. Thus suppose we have an ensemble of single spins with spin I, acted on by a static external field. Then ?to is the Hamiltonian of a single spin:
In the high-temperature approximation we can expand the exponenlial, keeping only the first tenns. By utilizing the fact that Tr {M z } = 0, we have
(M,)
=
~ T'{M,(,- ~+)}
~ ~ Tr{ (~~h::oI;) } Now, in the high-lemperanlre limit, Z = 2I + I. Since Tr {4} = we get
(M z ) =
~~fI2 I(l + 1) H 3kT
(5.93)
j I(I + 1)(21 + 1), (5.94)
0
which we recognize as Curie's law for the magnetization. The density matrix gives, therefore, a convenient and compaCI way of computing thennal equilibrium properties of a syslem. One situation commonly encountered is that of a Hamiltonian consisling of a large time-independent interaction 'Ho, and a much smaller but time-dependent tem H,(t). The equation of motion of the density matrix is then
de i dt = !i[e, 'Ho + Hd
(5.95)
If HI were zero, the SolUlion of (5.95) would be e(t) = e -(i/" )1io/ e(O)e(i/h)1i oi
(5.96)
Let us then define a quantity e* (the siar does not mean complex conjugate) by the equation e(t) = e -(i/")1iot e* (t)e(ijh )"Ho/.
(5.97)
If, in fact, HI were zero, comparison of (5.96) and (5.97) would show that e* would be a constant. (Note, moreover, Ihat at t = 0, e* and e are identical.) For small?tl, then, we should expect e* to change slowly in time. Substituting (5.97) into the left side of (5.95) gives us the differential equation obeyed by e*:
-~[?to, el + e- O/")1iotd: * e(ifh)1iot = ~(e, Ho + ?til t
(5.98)
We note thai the commutator of e with ?to can now be removed from both sides. Then, multiplying from the left by exp (il1l)'Hot) and from the right by exp (-(i/h)Hot), and defining
(5.91) We shall illustrate th~of the density matrix 10 compute the average value of the z-magnetization (1I1 z ). It is
(M z ) =Tr{Mze} =
162
~ Tr{M ze-"Ho/k1'}
(5.92)
Hi = ei1iot/h?tle-(i/h)"Ho/
(5.99)
we get, from (5.98),
de* i * * dt = hie, 1£,(')]
(5.100)
163
Equation (5.100) shows us, as we have already remarked, thai Ihe operator r/ would be constant in time if the perturbation ?i1 were sel equal to zero. The transformation of the operalor?il given by (5.99) is a canonicallTansformation, and Ihe new representation is termed the interaction representation. The relationship of U and (J. is illustrated by considering the expansion of the wave function tP in a form
tP = L:ane-(i/l)E.l un
We can make a closer approximation by an iteralion procedure, using (5.109) to get a bener value of U·(t l ), to put into the integrand of (5.108). Thus we find
·e(l) = ,'(0) +
,'(0) +
.
(5.101)
= ,'(0) +
o
(e(O), "H;(,")Jd," }. "H;(")] d"
0
(,(0). "H;(")]d"
o
n
instead of
2 t tl
+
(5.102)
tP=L:Cnu n
(nll?·lm) = e{i/Ii)(En -E".)t(nll?lm)
((,'(0). "H;(''')J. "Hi(")]d"
d'"
(5.110)
(5.103)
We could continue this iteration procedure. Since each iteration adds a term one power higher in the perturbation 'Hi, Ihe successive iterations are seen to consist of higher and higher perturbalion expansions in the interaction ?i I, For our purposes, we shall not go higher than the second. Actually, we shall find it most convenient to calculate the derivative of (J*. Taking Ihe derivative of (5.110) gives
Since (5.101) and (5.102) give the same tP, we must have
d,;,(1) = *["(0). "H;(I)] +
(5.104) (5.105) Comparison with (5.103) shows that ana~ :::::
(D JJ o 0
n
where the un's and En's are the eigenfunctions and eigenvalues of the Hamiltonian ?io. In the absence of ?iI, Ihe an's would then be constant in time. We shall show Ihat Ihe matrix ana~ is simply (nlu·lm). We note first that replacing I?(O) by U· in (5.83-85) gives that
(nil 1m)
Q.E.D.
(5.106)
There is likewise a simple relalionship between (nl1ii 1m) and (nl1i 1 lm). By an argumem quite identical to that of (5.83-85) we have (nj1iiJm) =
Ju~e(i/l)1tot1ite-(i/l)1totumdT
::::: e(ifA)(E"-E,,,)t(nl1i 1Im)
,'(I) = ,'(0)
+
(5.107)
u"
(5.100). By inte-
,
*J .
[,(I'), "H;(I')]dl'
(5.108)
o
This has not as yet produced a solution, since e·(t') in the integral is unknown. We can make an approximate solmion by replacing e"(t l ) by e"(O), its value at t = O. This gives us
.
'(') = ,'(0) +
,
i J(,'(0). "H;(,')Jdl' o
US J
((,'(0). "Hi(I')]. "Hj(l)]dl'
. (5.111)
o
It is important to note that (5.111) is entirely equivalent 10 ordinary timedependent perturbation theory carried to second order. However, instead of solving for the behaviors of a" and am, we are solving for the behavior of the products ana:"n which are more directly useful for calculating expectation val-
"es.
5.5 The Rotating Coordinate Transformation
.
Now we proceed to solve the equation of motion for grating from t = 0, we have
164
*J[{, *J" , *J
(5.109)
We saw in Chap. 2 that it is often convenient 10 go 10 the rotating frame when a system is acted on by alternating magnetic fields. We explored both a classical treatment and a quanlum mechanical treatment. In the laller we ITansformed the Schrtldinger equation. We now examine how to transform Ihe differential equation for the density matrix.. Let us consider first the case of an isolated spin acted on only by the static field Ho and a rotating field HI(t) given by (5.112) where we have already recognized the sense of rotation for a positive "'( in the negative sign in front of j. In the laboratory reference frame de i - = -(,"H - "H,)
d'
(5.113)
h
165
11. = -")'h[HoIz + HI (Iz cos wt - I y sin wt>]
(5.114)
Utilizing (2.55) we express this as
11. = -")'h[HoIz + Hle+ iwU , Ize- iwU ,]
where 17k expresses the chemical or Knight shifts. We omit a chemical shift correction from H I since HI <: Ho always. Defining
(5.115)
Defining an operator R as
(5.125) we now express 11. as a sum
R =: eiwll •
(5.116)
'H = 'Hz + 'H p + 'HI
with
(5.126)
we have
11. = -")'h(HoIz + HIRIzR- I )
so that
(5.111)
~: = itrr(HoIz + HI RlzR- I ) -
i")'(Holz + HI RIzR-1)e
(5.118)
Following the same reasoning as in Chap. 2. we now seek to eliminate the operators by defining a new variable eR
en
==
R-leR or
"a'
(5.119)
e=RenR-1
(5.120)
'H p =
k
. deR _I l(wIze- f!W1z) + R-;[lR
'HI = _")'hHleiwll'Ize-itNlI. Following the reasoning of Chap. 3, we keep only that portion of the spinspin coupling which commutes with 1iz. Thus. we express 'H il• as a sum of teons like the A, B. C• ...• F of (3.7), keeping only 1f!!k' the teons which commute with I z : J
(5.121)
11 =
Multiplying from the left by R- I , from the right by R. utilizing the fact that I z and R commute, and employing (5.119) and (5.120), we get
i
= h(eJ~11.eff -'Heffeu)
where
(5.122)
(5.123) is our old friend, the Hamiltonian of (2.63), which expresses the interaction of the spin with the effective magnetic field in the rotating reference frame. Let us now allow there to be more than one spin. allow the spins to interact with each other by lhe dipole-dipole coupling or by indirect spin-spin coupling through bonding electrons, and include Knight shifts and chemical shifts as well. Then, labeling individual spins by k or j, we have the Hamiltonian in the laboratory frame as
(5.124)
166
(5.128)
Thus we take as our total Hamiltonian 'H' in this perturbation sense
= -i,,),Ho«(!Iz - I z (!) -i")'H I (eRlzR-I - RIzR- I e)
deR
(5.127)
kJ
11f;,.I.)=0 .
Substituting (5.120) into (5.118). we get
dt
I
L")'h1ioUk I zk + '2 E'Hjk
,,),hHoIz + E ")'hHoUkIzk
,
-
+~
E1CJk - ")'hH,RIzR-'
(5.129)
j,k
We can now repeat the derivation of (5.122) above, noting that our new Hamiltonian differs from that of (5.114) by addition of the teons I
L; ,hHou,I., + 2 L;1I1j k
(5.130)
jk
all of which commute with R. (It IS Interesting to note that this expression through its 17k seems to require that we know the absolute chemical shift relative to a bare nucleus, i.e. a nucleus stripped of all its electrons. That viewpoint is wrong. All one needs do is pick a convenient reference compound for which one defines the chemical shift to be zero. ")' is then defined as being the ratio of the angular frequency to the static field for resonance for that reference compound. We discuss this topic further lit the end of this section). Thus we readily find
den &
i
I
I
= h(en'Jterr - 'Heff(!R)
where
(5.131)
167
'H~ff '" - ,MHo - w/-y)I, - 1hH I I r
I
+ "E1 IiH0t1 IJ,k+"2
L'H2 j
(5.139) (5.132)
k,j
Ir;
This new effective Hamiltonian reminds us by the presence of the tenns 'Hqlr; that if spins interact with one another in the laboratory frame, they do also i/the rotating frame. It also reminds us from the tenns involving UIr; that when there is more than one chemical shift, one cannot be simultaneously exactly on resonance for all nuclei. Since 'H~tr is independent of time, we can solve formally for en(t) in terms of its value at an earlier time (t '" 0) by eR(t) '" e -i'H~ffL/. e R(O)ei'H~ffl/lI
(5.133)
It is imponam to keep in mind that in general the interaction representation differs from the rotating coordinate transformation. Both may be used with a large time-independent Hamiltonian 'Ho and a time-dependem ?it(t). e- (e expressed in the interaction representation) is related to e by
e- (t) '" ei'Hol/h efJ)e -i11{Jtlh whereas
eR
is related to
(5.134)
and a new zero of chemical shifts
'171, '" Uk -
(j
(5.140)
1(z =
--I hHoI,
(5.141,)
1tp
L 7 'IHOUkI,1r;
(5.14Ib)
which makes
=
1
k
where in getting the second equation we drop terms such as 7iult. Clearly the fonn of (5.141) is the same as that of (5.127). We have simply redefined 7 to COlTeSpond to a different zero of chemical shifts. While (5.138) guarantees that the Zeeman term falls in the midst of the resonance lines, it seems to imply through (5.137) that we know the individual uk's, and go through some evaluation to get 7i. What is imponant to realize is that there is some arbitrariness in the division between 'Hz and 1tp which really corresponds to the arbitrariness in the choice of a zero of chemical shift.
e by
eR(t) '" e-K..>tI. e
(5.135)
5.6 Spin Echoes Using the Density Matrix
If we have the Hamiltonian of (5.114),
l
(t) '" e -K..>oll. e
(5.136)
These do become the same if we are exactly at resonance (wo '" w), but othelWise would differ. If the only time dependence in the problem is from HI, the rotating coordinate transformation eliminates it. The division of the Hamiltonian of (5.127) into ?i z and ?ip is nOt the only division one might think to make. For example, if one defines an average chemical shift for the system of N nuclei, 7i, by the equation 1 7i'" NLt1k
(5.137)
k
then one might prefer to have the Zeeman energy be the average value including the chemical shifts. We can do this by making a division into an 'HZ and ?il> defined as (5.138') ' 1 hH o(Uk - -U)I,k 'H' '" 'LJ p' k
, +"2I ",--?tij
(5.138b)
j,k
Note, however, that (5.138a) can be converted to (5.127) if we define a new 1, called
i,
168
Now that we have obtained the differential equation for the density matrix in the laboratory frame, in the interaction representation, and in the rotating frame, we tum to methods of solution. The rest of this chapter is devoted to that task. We will first discuss several problems involving pulse techniques, then we tum to use of the density matrix to discuss relaxation. As topics involving pulse techniques we shall first examine how to use the density matrix to derive the spin echo, giving us a chance to use this formalism to examine a problem whose physics we have now seen both classically and quantum mechanically (using wave functions). We then turn to some new physics which underlies one of the major techniques employed in NMR today, Fourier transform medlOds. To do this we first set the stage by deriving a result important to linear response theory, the theory underlying Fourier transform NMR. This result is the response of a system to a 8-function excitation. Anned with that result, we next go on to explain what is meant by Fourier transfonn NMR, and then prove the fundamental theorem which specifies what it measures. We then further develop a concrete appreciation of the density matrix by exploring what happens to the actual matrix elements under the action of an alternating magnetic field. The most general approach to solving the equation is to pick an appropriate family of wave funclions as a basis set, then conven the operator equation to a set of coupled differential equations between the matrix elements of e and ?i. Thus if we pick a basis set 10), where 0 stands for the set of eigenvalues which distinguish the states, we convert the equation 169
d(!
i
- = -«(!H - ?i(!) dt h
to
j;
d (aleld) = D(aleIP)(P1Hla') - (aIHIP>(Plela'» • p d,
H,
(5.142)
(5.143) I
In many cases there are then straightforward melhods of solution. Once one has
r------ T ----l
I
iI
obtained the (alula'rs, one can then calculate the expectation values of observ~ abies using (5.77),
I
I
"
t1
t2
t]
(M.> =T,{M.e) = L:(a'IM.la)(alela')
(5.144)
o
Fig. 5.4. III vs time, giving the definition of the times 0, tl, t2, and 13 which spc<:ify the beginning and end of the tr/2 pulse (I = O,td and of the tr pulse (t = t2,13)
0',0"
The weakness of Ihis appro,\lch is that it may fail to reveal important physical insights available from other methods. The "other methods" are based on manipulating operators, and are analogous to the quantum mechanical derivation of the echo carried out in Chap. 2. We now tum to several illustrations of the use of the operator approach. Later we return to use of (5.143). We begin with a derivation of the spin echo of a group of non interacting spins using a (rr!2,7r) pulse sequence. We deal with this problem by first transfonning to the rotating frame so that our basic equation is given by (5.123)
d(!R
--;It
As in Chap. 2, we approximate this as (5.146)
1i"err = -1'hH I T;I; Then, utilizing (5.83), we write I? n(t I) = e iWI tl f" I? n(O+)e -iWl 11 I"
= eiW1 tl f z I?n(O-)e-iW1 tIl"
(5.148)
WI = 1'HI
i
= "h«(! llHeff - 'Herru n)
(5.147)
where
Likewise, during the interval tl to t2, where HI is zero, we have
As in Chap. 2, we assume that the static field is inhomogeneous with a distribution function p(ho)dho for the N spins, where dN, the number of spins whose resonance lies between ho and ho + dho, is given by (2.119):
fm(t2) = e i ,.Il o(t2 -( 1) I, I? n(t l)e -i,.lt o(t2 -11 )1.
Defining [as in (2.122)] X(8) =
dN = Np(}to)dho
(5.149)
e ifJl"
(5.150,)
and where ho is given by (2.118):
(5. 150b)
ho = H o -wh We take the distribution function to be symmetrical about a center value (lto = 0) and we assume w is to be at resonance when ho = O. Since we have already treated this problem in Chap. 2 using wave functions, we will employ the same notation (Fig. 5.4). At t = 0-, the density matrix has its initial value un(O-) which represents thennal equilibrium in the laboratory frame. At t = 0, HI is turned on. Although there is a discontinuity in 1i"eff at t = 0, the right-hand side of (5.123) remains finite. Therefore, integrating (5.123) with respect to time from t = 0- to t = 0+, we find
selecting WI tl = 7r/2 and WI (tJ - t2) = 7r for the rr/2 and rr pulses, and choosing times t later than t3, we get (!n(t, ho) for those spins off resonance by ho as I?n(t, 110) = T(t - 7, ho)X(rr)T(T, h o)X(rr/2)l?n(O-) x X-I(rr!2)T-I(7,ho)X-I(rr)T-I(t - r,ho)
(5.151)
Following our procedure from Chap. 2, we insert X(rr)X- I (7r) = I to the left of T(t - T, ho) and utilize (2.134) to get X(7r)X- t (1l")T(t -
T,
ho)X(rr) = X(rr)T-1(i - 7, ho)
(5.152)
(5.145) A similar argument holds of the other discontinuities. During the time interval from t = 0 to t = t I or from t2 to tJ, the Hamiltonian is given by (2.117)
1teff = -1'1i(h oI z + H IIz) 170
I?n(t, ho) = X(7r)T- I (t - T, ho)T(T, ho)X(rr/2)l?n(O-) x X- 1(7r/2)T- 1(T, ho)T(t - T, ho)X-I(rr)
Thus, when t -
T
(5.153)
= 7, or t = 27, 171
(l/l(2r, 110) '" X(7r)X( 7r(l.)e n(O-)X- 1(7r(l.)X- 1(7r)
(5.154<1)
= X(3./2)eR(0-)X-'(3./2l
(5.154b)
= X(-'/2leR(0-)X- 1(-./2)
(5.154c)
where the last step utilizes the fact that X(2Jf) '" X(O) when X operates on fl [contrast this with (2.78)]. Equation (5.154) shows that when t '" 2r, fl is independent of Ito, hence we have an echo. Moreover, flR(2r) is the same as we would have if instead of applying two pulses spaced apan by T, we had applied a single pulse which produced a -1f/2 rotation. We now proceed to calculate the magnetization signal at the echo peak. Since (;lR(O-) corresponds to thennal equilibrium in the laboratory frame, we have, using (5. tl9) and (5.90),
Tr{G} ",Tr{I+X(-7r!2)(1 + -Yli:'~lz)X-I(_7r/2)}
'" ..!.-e-x....II. e -'Ho/kTeilollf• Z Since 'Ho '" -'YhHolz, we get
~eY"l1o/./kT
(5.155)
T,(I+X(-./2lX-'(-./2») =T,(r+) =0
_ I ('YhHolz ) (;lR(O )"'(21+1) 1+ kT + ...
{ h)'N H
,H, {e/l(t)I.) ,H, {e/l(t)Iy )
(5.158)
Combining we get
(M.(t) + i(My(t) = (M+(t» =
,I, Y, {e/lI+)
(5.159)
Since both (M~(t» and (My(t» are real, we can get them both readily by taking the real and the imaginary pans of (5.159). We utilize (5.154), (5.156), and (5.157) to calculate (.M~(2r» and (My(2r»
Jp(l'o)dh"h (r+ 1I<2T, ho») ,hN J + P(ho)d11o {G}
172
T,
I)
1
Tr
(5.162)
Since Tr {l>;Ig }
'"
~;;~~
:;y,
(r+(-ly))
(5.163)
0 and Tr{l;} '" (21 + 1)1([ + 1)13, we get
N-y 2 h 2 I([ + 1) . . 3kT Ho' = -IXoHo
.
(5.164)
(5.157)
"0.
'" (21
(5.161)
Now X(-1(/2) '" X-I (1f!2). Hence, using (2.124) (X- t ('lf!2)lzX(Jr!2) '" -llIl we get
(M+(2T» = -
(Note that since the spins do not interact, we can work with the Hamiltonian of one spin. However, we eventually include the effect of all N spins when we This situation is to be contrasted with that of SecI.S.7 where integrate over the Hamiltonian is that of N coupled spins). We have operators M>;(", -ylJi>;) and My corresponding to magnetization along the x·axis (the HI direction) and the y-axis in the rotating frame given by
(M+(2T» = N
.
The physical significance of this result is that at infinite temperature, where only the a term remains, the thermal equilibrium magnetization vanishes. Therefore, no pulse sequence can produce a signal. We will encounter a similar situation in all pulse sequences stnning from thermal equilibrium. The b term gives
In the high temperature approximation, Z '" 21 + I, giving
(M.(t) = (My(t» =
,
"
In the trace, the term labeled a is readily shown to be zero:
(M+(2T» = (5.156)
Z
(5.160b)
'--v-'
(M+(2T» = (;I + I) k;Tr{r+ X(-./2lI.X- (-./2))
(;lR(O-) '" e- iloltl • e(O-)eilo'll.
(;lR(O-) '"
where
(5.160<1)
The -i shows that at t '" 2T the magnetization has zero component along the x-directjon (i.e. HI), that its magnitude is the same as the thermal equilibrium magnetization, xoHo, and Ihat it points along the negative y-direction. The approach we have described can be extended readily to deal Wilh arbitrary sequences of pulses. Let us consider that at t '" 0- the system is in thennal equilibrium with density matrix (;In(O-) '" e(0-), as in (5.152). Stnning at t '" 0 the Hamiltonian in the rotating frame takes on a value 'HA for a time interval tAo then jumps suddenly to 'Hn for a time lB' and so on. In the example above tA '" t]
11.A '" --y1i(holz + H l l%)
tB",t2- t l
1{B '" -"'f1ihOlz
tc'"
He'" --yli(IIoIz + J[ll~) 'HI) '" -,liholz
t3 - t2
tD"'t -tJ
(5.165)
Corresponding to t A and 'H A , there is a time development operator Til given by (5.166)
Til '" exp(-i7tAtA/li) Similar equations hold for the succeeding inlervals. Thus, we get en(tz) '" TZTy ... TBT"{;l(O-)T; I Tjj I ... TylT I
Z
(5.167) 173
We chose to neglect all tenns in the Hamihonian but the tenn involving HI when HI is on. That approximation is often useful (as illustrated above) but would not be allowable if one were trying to understand what happens when HI is comparable to the line width. The general fonnula (5.167) does nOI necessarily include this approximation, but the approximation is readily introduced if desired.
Now lhat we have looked at a familiar problem with the density matrix fonnalism. we tum to a problem we have not treated quantum mechanically before. the response of the system to a o-function. This topic is imponam in pan because, as we saw in Chap. 2. if we know this response we can use it to calculate both I and I'. But it is also important because it underlies the theory of Fourier transfonn NMR which we take up in the next section. We therefore stan by recalling (2.156): ~
J,
~
J,
mer) cos (wr)dr
x" =
mer) sin (wr)dr
(2.156)
where mer) is the response at time r to a delta-function applied at r = O. We now tum to the calculation of mer). To do this we must first senle on the Hamiltonian. We take a Hamiltonian which has substantial practical utility. that of a group of N identical spins. coupled to each other with dipolar. pseudooipolar, and pseudoexchange interactions, experiencing chemical and Knight shifts. A more general Hamiltonian would include different nuclear species. quadrupole coupling. and relaxatiOll phenomena to a thennal reservoir. With the approximation above we will then have a Hamiltonian
'H = -
2:: ,hH,I,k(l k
1
Uk) + 1: 2::'H;j i,i
2:: ,hC'(t)Iz
(5.168)
k
which is similar to that of (5.123) except for the replacement of the HI driving teno with a magnetic field CE(t). The parameter C is used to define the size of the driving tenn. This is a Hamiltonian in the laboratory frame. In (5.168), the 8·futlction field has been applied along the x-direction in the laboratory. It will produce magnetization in general with components along the laooratory X-, y., and z-directions. Thus, in general such a drive produces a magnetization in the a'·direction for a o-function in the Q-direction. This fact requires that we label m(r) by subscripts a and 0'1 just as we label the X by X(I"(l' on p.58. Thus, let 111(1"(1'(1") be the magnetization in the a'·direction pf<xluced by a 6-function field applied in the a-direction. Equation (2.93a) gives X~", and X~",· The meaning of the constant C can be seen by considering the imegral of the driving field CE(t): 174
JC'(t)d'
,-
= C
(5.169)
But, 'by definition, m(t) is the response when this integral is unity, i.e. when C = I. or. alternatively, if M(r) is the response when C 4- I, meT) = M(T)/C
5.7 The Response to a 6-Function
X' =
,+
(5.170)
.
While in principle one can set C = I, in the process one conceals some units. and thus complicates the use of dimensional arguments to check the correctness of fonnulas. We will therefore keep C, realizing that eventually we will utilize (5.170). Indeed. to keep the units explicit. we will replace C by a product H TfJ:
Hro == C
(5.171)
where one can think of H as the height and 1"0 the duration of a rectangular pulse which. in the limit that rO goes to zero, represents the 6-function. To find mu(r) we need to solve the density matrix with the Hamiltonian of (5.168). with the initial condition that e(O-) represents thennal equilibrium. At t = 0, the driving teon dominates. since the 6-function is the biggest teon. At all ocher times. only the Zeeman and spin-spin teons are present Thus, if t > O. we have a time-independent Hamiltonian. We will treat it in the same spirit as the Hamiltonian of (5.126) in which we recognize that the spin-spin coupling lenns are perturbations, SO (hat it is appropriate to drop the teons in the spin-spin coupling which do not commute with the Zeeman coupling. We thus have as our Hamiltonian 'H.
=
-ihHol~ + L'Yl~Hou~JJ~1e + ~ L,'H.11c -
'YhlzHroo(t)
, (5.172)
i,k
Ie
where 'H1k is the ponion of 'H.jk which commutes with the mean Zeeman energy
II.. 'H1j l = 0
(5.173)
It is convenient, as in (5.126), to define three quantities
'Hz == 'H.p ==
'YhHoI~
L'YhHou~kI~k +~ "L'HJk
(5.174)
i,k
Ie
H6 = -ilil",HrOO(t) Note from (5.174) that 'H.z and 'Hp commute
['HI'. 'Hz] = 0
.
(5.175)
'Hz is the Zeeman energy, apart from chemical and Knight shift differences. Hp is a much smaller tenn ("P" for perturbation) which gives rise to splittings and 175
line widths. For t > 0, since we have then 'H6 = 0, we have a lime-independenl Hamiltonian 'Hz + 'Hr. Then, using (5.95), for t > 0,
e(t) = .. p (-i(Hz + Hp)t/IiJe(O+)exp(i(Hz + Hp)t/Ii)
.
therefore
(5.185)
(5.176)
To find e(O+)we must solve for the time development under the action of the C-function. Since we seek the linear response, it is appropriate to treat 'H6 as a perturbation, keeping only the fim-order lenn. We therefore turn to (5.11 I), using 'H6 for 'HI> keeping only the first lenn on the right. In this equation
e'(.) = exp (-i(Hz + Hp)t/Ii)e(t)exp (;(Hz + Hp)t/Ii) 'H (t) = exp( -i('Hz + 'Hp)t/h)'H6(t)exP (i('Hz + 'Hp)tlh)
(5.177)
de'(t) = .'.[e'(O-)H;(t) - H;(l)e'(O-)J
(5.178)
6
Then
dt
(5.186) The first tenn, linear in I z , corresponds to a thennal equilibrium magnetization. Therefore, when put in (5.176), it will not lead to any transverse magnetization. To get mn(t), therefore, we keep only the tenn involving I y , utilizing (5.170) (H Tl)mxx(t) = (Mx(t»
h
=,/oT,{Ixe('»
Integrating from t = 0- to t = 0+, we gel
,/oHo
= ,liT, { IxT(1) Z(oo)kT,(Hro)IyT
0+
e'(O+) - e'(O-)
=.'./0 J
[T(I)e(0-)H,(t)T- 1(.)
0'
0-
(5.179)
- T(t)H,(t)e(0-)T-1(t)]dt with
(5.180.)
T := exp (-i('Hz + 1ip)tlh)
and (for future use) Tp := exp (-i'Hpt/li)
.
(5. I 80b)
But over the zero time interval we can neglect the time dependence of everything except the c-function giving
.(0+) = e'(O-) - *,/oHTo[e(O-)Ix - Ixe(O-)J
(5.181)
and indeed
(5.182) Now e(O-) corresponds to Ihennal equilibrium, hence is given by I
1
e(0-) = Z ..p[-(Hz+Hp)/kT] '" Zexp(-1iz/kT)
(5.183)
In the high temperature approximation this gives
(I
+ ,/oHoI,) kT
(5.184)
Following the argument after (5.158), we drop the first tenn in the parenthesis since it contributes zero magnetization, getting 176
(I)
.
}
(5.187) (5.188)
A useful variant on (5.188) can be obtained utilizing the fact that 'Hz and 'Hp commute, and expressing Tz as exp (iwotlz), with wo = "tHo. Then
Tr{IzTIIIT- 1 }
Tz := exp (-i'Hztlh)
e(0-) = _1_ Z(oo)
"'I 2h2 Ho -1 mn(t) = Z(oo)kT "'I Tr {IzT(t)IyT (t)}
-I
= Tr {Til IzTZTpIyTpl} = Tr {e- lwotl• IzeifNOLf'TpIyTpl} = Tr {IzTpJyTp I} cos wot + Tr {Jy T p l yT p t} sin wot
This equation is closely related
10
(5.189)
(B.18) in Appendix B. At t = 0+, Tp =
Tj;l = 1 so that the coefficient of the cos wot tenn vanishes (Tr {IzIy } = 0), and the coefficient of the sin wot term is Tr {J;}, which is nonzero. For times shon compared to h/11ipl, where by l1-lpl we refer to the magnitude of a typical nonzero matri", element of 'Hr, we can slill appro"'imate Tp as 1. Consequently, the magnetization nlzz(t) will start as sin wot, a signal oscillating at the Lannor frequency. Since 'Hp gives rise to spectral width, over times of the order of the inverse of the spectral width, Tp will differ significantly from I, and we may e"'pect both traces to contribute. In Appendix B, Eq. (B.18), we show that if 'Hp contains only spin-spin terms and 7 is chosen such that the Uk'S are all zero, the first trace always vanishes. The effect of the operators Tp is to modulate the coefficients of the oscillations at Wo over times of the order of the inverse of the spectral frequency width, L1w. 177
Thus, for times short compared to the inverse of the spectral frequency width, ~, we may write mzz(t) = 7
721i2 Tr {12} Z(oo)kT Ho sin wot
with
t
<: l/L1w
•
It is straightforward to show that the term a for a many-spin system is the static susceptibility Xo. Hence
for
t
(5.191)
This result has a simple classical significance (Fig. 5.5).
[cos (woOTr {Io;Trl yT p1}
=
J
mu(T) cos (wT)dT
o
X~:I:"(w) =
y
,
....ig.55. A magnetization, Alo = xoJ1" initially along the :--direction is rotated through II. small angle dB = "'I J1 TO by a field J1, of duration TO, along the laboralory :I:"-direction. The rC:$Ulting magnetization has .. component Mod6 along the laboratory ",-direction. At later times, it precesses about z, producing .. nonzero component along the :I:"-axis which oscillates all sin Io'(Jt
(5.192)
giving My = XoJIO(7HTo) = HToxowo
(5.193)
This magnetization precesses at Wo, in the left-handed sense, giving a sin wot dependence to M:I:"' lllUs, utilizing (5.170), mo;o;(t) will be given as
(5.194)
for times short compared to the dephasing effects arising from chemical and Knight shift differences and spin-spin couplings. We see, therefore, that (5.188) and (5.189) have simple physical significance. In fact, we can usefully rewrite them as 178
=
J
(2.156)
mu(T) sin (wT)dT
o
Considering the 6-function to be a field in the x-direction of strength H>- Ho, on for a time 7"0 < I/wo, we see that the field tilts the thermal equilibrium magnetization xoHo to have a component in the positive laboratory y-direction. Since we are seeking a response (Le. an x-y magnetization plane) linear in the driving tenn, H, we keep the angle of rotation, L18, small
mu(O = xowo sin wot
(5.195)
While (5.195) looks relatively simple, involving just a few traces, its actual evaluation may be enonnously difficult since the equation contains all the content of the theory of line shapes! Nevenheless, the equation gives a compact statement for attack by the various fonnal methods which have been developed to deal with the line-shape problem. In principle, this mu(t) will yield both the absorption and dispersion spectra utilizing (2.156)
x~z(w) =
z
X{w~}
Tr l y
+sin (woOTr {IyTplyTj;I}]
(5.190)
~
mu(t)=xowo sinwot
mzz(t) =
5.8 The Response to a ,,/2 Pulse: Fourier Transform NMR We tum now to the topic of Fourier transfonn NMR, a method of doing magnetic resonance which is today all pervasive in the field of high resolution liquid spectra, and which fonns the basis for one of the most important other developments in NMR, so-called two-dimensional Fourier transfonn NMR (which we take up in Chap. 7). In 1957, Lowe and Norberg [5.6] discovered an imponant theoretical result which was confinned experimentally by them in conjunction with Bruce [5.7]. Lowe and Norberg showed theoretically that in solids with dipolar broadening, the Fourier transfonn of the free induction deCa'yfoliowing a -rr/2 pulsegave the shape of the absorption line. They verified their result experimentally with a single crystal of CaF2 by comparing their experimenlal free induction decays with the Fourier transfonn of Bruce's steady-state absorption data taken on the same crystal. In 1966, Ernst and Amfer.fon [5.8] pointed out that there were substantial experimental advantages to the use of pulses over steady-state meth<X1s when dealing with complex spectra. (We discuss advllntagcs later in this section.) Their method, which consists of first recording the free-induction signal, then laking ils Fourier transfonn, is referred to as Fourier transfonn NMR. The method allows on~ to obtain both the absorption spectrum and the dispe-;;ion spectrum'"":", ~how ttlatthe cosine transfonn of the free induction decay gives one whereas the sine transfonn gives the other. (See also the book by Ernst et al. [5.9].)
TheY
179
[ These remarks are reminiscenl of linear response theory discussed in Sect. 2.11 where we found thut X' and i l are given by cosine and sine transforms of the function m(T), (2.156):
J
~
x' =
~
m(T) cos (wi)dT
X" =
o
J
m(T) sin (wT)dT
(5.196)
~n
_1_(1
(0-)= (0-)= U Z(oo)
1C = 'Hz +'Hp +1-{1
with
(5.197)
(5.198)
(5.199) (5.200) expressing the fact that the spin is acted on by a rotating magnetic field which is to generate a 7f(l pulse. During the pulse, we go to the rotating frame where UJl is described by (5.131):
i i ,
= "h(un'H eff - 'HeffUIl)
(5.201)
with 'H~rr given by (5.132) and UIl by (5.116) and (5.119). While the H t is on, we take it to be so large that we can neglect everything else in 'HI giving
'H' = -"(fi.HtI~
.
hH 1 o ,) kT
(5.204)
1
-I)
un(t,) = Z(oo) (lhHO 1+ kT X(wltt>I~X (Wltl)
(5.205)
where X(D)
== e+ i8Jz
(5.206)
Now, ulilizing (2.55),
X(8)I z X-'(8) = I y sin 8+ I~ cos 8
so that UR(tl) =
Z(~) (I + 1::0 [Iv sin Witt + I z cos Wtttl)
(5.207)
(5.208)
The first term on the right will nOt contribute 10 the magnetization, following the argument of (5.160) and (5.161) (i.e. it is allihal remains as T-+oo). Moreover, the term involving I z is, apart from a constant of proponionality, the same as the temperature-dependent contribution to a (! R corresponding to thermal equilibrium. Thus, it contributes nolhing to transverse magnetization. (The validity of neglecting the tenns involving 1 and I z can be verified by direct calculation of (Mz(t)) and (M!I(t)).] We are left, then, with I 1liHo. UR(tl) = Z(oo) Sill (wltdIv
---;;r
(5.209)
Therefore, utilizing (5.120),
e(tl) = RUR(tI)R- I I l'IHo. -I = Z(oo) kT ,m (w",)[R(t,)I,R ('1))
(5.210)
where R is given by (5.200): ROI) = ei""lt J•
(5.211)
To gel (!(t) for t> tl' we introduce the operators T, Tz, and Tp defined as in (5.180),
(5.202)
If the HI is turned on suddenly at t = 0 and turned off at t = tit we have, then [os in (5.147)1. UR(tI) = e+iw]It r. un(O+)e-iwtltJ" = e+iW ]lt 1"un(O-)e-i""t 1tl"
+7
whence, as in (5.150a) and (5.151)
•
where m(i) is the response of Ihe syslem 10 a S-function which we treated in Sect. 5.6. We therefore demonstrale that in the linear response regime al high temperatures, a 1f(l pulse (in faCI, any pulse) is equivalent to a S-funclion excitation, hence Ihat the transforms of the free induction signal give ;( and X". Let us consider, then, Ihe Hamiltonian of (5.124) describing a group of N coupled spins, allowing there 10 be both chemical and Knight shifts. We then have a Hamiltonian 'HI given by keeping only Ihat part of the spin-spin coupling which commutes with the Zeeman coupling as in (5.172)
dUll dt
I
with WI = "(HI. Taking UR(O-) to represenl thermal equilibrium, we have, utilizing (5.155-157) generalized for the Nobody Hamiltonian as in (5.184),
(5.203)
Tz(t) = exp (-i'H.ztl1l) = exp (iwotIz) Tp(t) = exp (-i'H.pt/h)
T(t) = exp (-i(1ip + 1iz)tlh) = TzTp '" TpTz
(5.212)
(Do not confuse the operator T(t) with the temperature T!) 18.
181
Thus, since 1tz + 7-ip arc independent of time, we get from (5.83)
Utilizing (5.196) and (5.219) with Hro = I, we have then
=
e(l) = T(t - tt)e(tl)T-I(t - til =
-yIiHo
Z(oo)kT
sin (Wltl)T(t _ tl)R(t\)lyR-I(tl)T-IU - t\)
T(t - tl)R(tl)
= exp (iwotI:) exp (-i1tpt/fi) exp (iHptdh)exp [i(w - WO)t1 1:] (5.214)
If, now, we neglect the effect of Hp and (w-wo)t I during t\ we can approximate Trl(tl)ei(""-~)tl/. ~ I
(5.223b)
o
= exp [iwo(t - t 1)1:] exp (-i7-ip(t - t I )/1i.) exp (iwt I [:)
p
.'0 J(Mz(r»o sin (wr)dr sm
=
Xl/(w) =
= T(t)T I(tl) exp (j(w - WO)tl 1:)
(5.223.)
Sin 0 0
..
(5.213)
Now
(5.215)
getting
Equation (5.223b) includes the result of Lowe and Norberg (they did not include chemical or Knight shifts). Equation (5.223) includes the result of Ernst and Anderson that the cosine and sine transfonns yield respectively the dispersion and the absorption. [It is useful to keep in mind that the tenns "cosine" and "sine" imply a particular phase convcntion. Ours is defined by (5.221).] The validity of (5.218) rests on the approximation that during the pulse we can neglect the effect of Hp and (w - woH I. Since HI' leads to a spectral width, Llw, this approximation is equivalent to
(5.216) Therefore, the expectation value of the transverse magnetization (Mz ), following (5.158), is
(M.(t)
=
(5.217)
This result is identical to the result of (5.187) except that sin WI tl replaces -yHro. We introduce the notation
(Mz(t»o = -y
2,.2H
t 0 sin (O)Tr{IzTlyT-I} Z(oo)kT
(5.218)
the magnetization following a pulse with wltl
(5.219)
== 0
Then we can write sinO
(Mz(t))e = 7n zz (t)-H ,
TO
(5.220)
Now, (Mz(t»e is just the signal we would observe from a pulse in which -yHltl = 0, produced by a rotating field which at t = 0 lay along Ihe x-axis
according to the convention of (5.114). This is also the rotating field generated by a linearly polarized field along the x-axis given by Hz(t) = Hzo cos wt
(5.221)
where, following (2.83),
Hzo 182
= 2H I
"YHI »Llw
and
(w - WO)tl «I
and
"'fHI» Iw - wol
or
(5.224.) (5.224b)
The classical derivation of the b-function response given in (5.192-194) involved assuming that a magnetic field H satisfy
,1m {g(t)I.}
2h2H O Sin . (Wltl )Tr (I z TIy T- 1) = -y Z(oo)kT
j
(Mz(r»o cos (wr)dr
i(w) = ."1
(5.222)
H»Ho
(5.225)
where the field H lasts for a time ro. This condition signifies that the magnetization is rotated during a time less than a Lannor period in H o. This appears to be a much stricter condition than (5.224b). The reason one can use an HI «Ho is that HI is rolating at a frequency w which is nearly the same frequency as the free precession frequency WOo Thus, while HI is on, the total phase angle about the Ho axis which the spin advances is wtl. If a l5-function had kicked the spin at t = 0, the subsequent precession angle in the same time tl would have been wotl' But since I(wo - w)tll and Llwtl «I we can take wOtl = wtl' Thus, the HI builds up the rransverse magnetization over the time it> but does not in the process inrroduce a phase error relative to a "zero time", or 6-function, creation of transverse magnetization followed by free precession for t I. We note also that this argument explains why there is nothing magical about exciting the free induction decay wilh a 0 = 7T/2 pulse. Equation (5.221) shows that any angle of 0 will do. [The fonnula (5.221) appears to blow up for sin 0 = 0, which simply reflects the vanishing of (Mz(t»o for that case]. There are several very imponant advantages to the Fourier [Tansfonn approach. The first is that since one is recording the NMR signal when the HI is zero, one eliminates noise from the oscillator which drives the HI. For steadystate apparatus, the oscillator noise is reduced by bridge balance, but typically such an approach reduces noise which arises from amplitude modulations of HI but does not reduce noise which arises from frequency modulation. 183
If one is dealing with a complex spectrum which has many absorption lines well separated from one another, as in many high resolution spectra of liquids, there is an additional advantage to the Fourier transfonn method. A steady-state search necessarily spends much of its time looking between the resonance lines of such a spectrum. The search procedure is thus inefficient. By pulse excitation, all the spins are induced to broadcast their presence at their characteristic frequencies, thus one immediately gains in dam collection by eliminating the dead time of a steady-state search. Of course, one must use a broader band system, since the steady-state system need only have the band width necessary to examine a single line whereas the pulse system needs a band width large enough to include the whole spectrum. However, since the noise voltage goes as the square root of the band width, one still is better off with the pulsed approach. An important point is that the Fourier transfonn method is the basis for the enonnously powerful two-dimensional Fourier transfonn techniques which have revolutionized NMR. II is useful at this point to make several remarks about instrumentalion. The signal M:c(r) consists in general of a sum of oscillating signals (corresponding to each resonance frequency which has been excited) which decay as a result of line broadening and relaxation phenomena. In general, all of the spread in oscillations is small compared [Q the Larmor frequency. To record these frequencies takes very high frequency response. It is customary to use a technique called "mixing", which in essence translates the fTequency of oscillation while preserving phase relationships. We will not go into the details of how mixers work, but essentially they work as shown in Fig. 5.6. The mixer has three pairs of terminals, one at which the signal is applied, one to which a reference voltage is applied, and one which is the output, the "mixed" signal.
Mixer Signal voltage input o---+~
Mixer signal output
If Vsig(t) = Vs
cos (wst
+,p)
= Vs(cos ¢ cos wst - sin r/J sin wst)
(5.226,) (5.226b)
Vrcr
Then, Vout = Vs cos [(ws - wn)t + ,p
- 0]
(5.227)
In expressing the reference, we have intnxluced a phase angle 0 to represent the fact that the phase of the reference can be set by the experimenter. Suppose 0 = 0, then (5.228) At firs! sight, this expression appears to be equivalem 10 (5.226a). However, there is an ambiguity since the cosine is unchanged when the sign of its argument is changed. Thus, one does not know whether wn < Ws or ws < Wit from recording (5.227) alone. To resolve the ambiguity, we record a second signal with (} = -1r/2. If 0 = -1r/2 Vout = Vs cos [(ws - wn)t + (
+,pl
(5.229') (5.229b)
Together, (5.228) and (5.229) enable us to determine both the sign of (wS - wn), and the phase angle ,p. Recording both such signals is called quadrature detection. It requires two mixers, whose reference signals differ by rr/2 in phase, and the sels of recording apparatus to record the output signals from both mixers. In the example above we took the reference phases to be "0" and "-1rI2". To apply (5.196) we must recognize that there is an absolme meaning to the phase since we took the rf driving magnetic field to vary as cos wt in our definition of Xl and XII [see, for example, (2.92), (2.95) and (5.198)]. I{ is best, then, to consider thal the two reference phases are (5.230)
Reference voltage input "'ig.5.6. Schematic diagram of a mixer. The radiofrequency magnetic resonance signal is fed inlo the signal voltage input terminats; a referenee voltage whose phase can be adjusted is fed into the reference volLage input. The voltage at the mixer signal output goes to recor
184
where 0 1 is unknown until one goes through some procedure to set it. Using 01> we then have from the first mixer Voul = Vs
cos [(ws - wn)t + ,p
-
Oil
(5.231')
+,p - Oil
(5.23Ib)
and from the second mixer Vout = Vs sin [(WS - wn)t
185
These two outputs together enable one to detennine Ws and ; - 8 t . If one has a procedure to detennine 81 in absolute tenns so that one can set it to zero, the cos output gives the dispersion signal and the sin output gives the absorption signal.
E p = (pl1tzIJI) = --yhHoJt = -1""'OIJ
We use these as basis states to express
e,
.
(5.241)
(5.235), in matrix fonn:
A(I I dt pen I') P = '"i .... L;:«P Ien III Ii )(/1 /I j'Hefflp) 5.9 The Density Matrix of a Two·Level System
"
- (/tl'Herrlpll)(Jt"lenll/» We have seen examples of the use of operator methods in conjunction with the density matrix to calculate the results of a sequence of pulses. These methods are extremely powerful and help us see what is happening in physical tenns. However, there is also great insight to be found in actually following what is happening to the individual matrix elements of the density matrix for an example which we already understand thoroughly, the effect of an altemating field on a spin! system. [The spirit is that of (5.143).] The Hamiltonian in the laboratory system for this problem is then
'H = -"(hHoI: - "'(1tH I (Iz cos wt - I,I sin wt)
(5.232)
where we assume a positive "( and a rotating field. As in (5.115), we write 'H. = -"(h(HoI: + e'loItl z Ize-iloltlz) Defining, as in (5.116),
R = e ilolU•
(5.234)
we then transfonn to the rotating coordinme frame to have (5.122)
deR
"""dt
i = ",(eR'H.err -'HerreR)
(5.235)
with
'Herr = -"(1t[(Ho - wl"'()I: + HtIzl
and
(5.236.) (5.236b)
We take Ho = wh, exact resonance, giving 'Heir = -1twl I z
with
,
'86
p=±!
(5.242)
and so forth for the other matrix elements. Recognizing that
(pll,I"') = 0
(5.243)
unless p 1= p', we then get four equations such as
d"++ . [ --:it = -lw\ g+_<-Il,I+) - (+ll,l-lg-+]
de++
dt de__
iWI
= -T(e+- - e-+)
de+_
iWI
de_+
iw)
--:it = -2(g++ --:it = -2(g-- -
e+-)
(5.245b)
g--)
(5.245c)
g++)
(5.245<1)
We note immediately that if HI = 0, all elements of eR become independent of time. They change with time only while the pulse is on. Adding (5.245a) and (5.245b) we get
e++(t) + e--(t) = const = "++(0) + g__ (O)
(5.240)
(5.245.)
iWI
d t = -T(e-+ -
(5.238)
(5.239)
(5.244)
t, so that
de++ + de __ = 0 dt dt '
hence are given by eigenfunctions and eigenvalues of I::
1,lp) = pip)
(1Ignl- ,) = g+-
(5.237)
The energy levels and eigenstates in the laboratory system in the absence of HI are those of the Zeeman Hamiltonian 'H.z = -"(!tHo I:
A somewhat more compact natalion is useful, defined as
But (+IIzl-) =
(5.233)
.
or
(5.246)
(5.247)
a stalement of the conservation of the tOlal probability of finding the spin in either the up or down spin states. In a similar way, adding (5.245c) and (5.245d) gives
e+-(t) + e-+(t) = const = "+-(0) + g_+(O)
.
(5.248) '87
If we take the differences instead of the sums of the two pairs of equations, we get
~(E?++ -
l?--) = -iwt(l?+_ -
d'
(5.255b) Expressing these in matrix form, we get
l?~+)
(M+(t)) = 1"
(5.249)
L:(,.len(')I,/)(P'Ir+I,.)
(5.256)
(M,(I)} = 1/' L:("len(I)I,I)(/II,I") These equations state that the off-diagonal marrix elements are "fed" by the action of HI on the population difference l?++ - l?-_. By taking the lime derivatives of (5.249), we get
Jl
2
Jl
2
dtZ(£,++-l?--)+wl(l?++-l?--)=O
/,,/,' which give, on utilizing the explicit matrix elements of 1+ and I:.
(M+('» = 1"e-+(t)
(5.257a)
(5.250)
(5.257b)
(5.251)
These equations show immedialely that the existence of transverse magnetization depends on the exislence of nonzero off-diagonal elements of l?R' Moreover, the z-magnetization is proportional to the difference in the diagonal elements of l?R. hence it will vanish unless £,++(0 j. l?--(t). The significance of (5.252) and (5.253) is clarified by considering e(O) to correspond 10 thermal equilibrium, in which case the off-diagonal terms l?+-(O) and l?_+(0) vanish. Note., incidentally that (5.73) requires that£,+_(t) and £,-+(t) are complex conjugates of one another. TIlen we get
Therefore,
l?++(t) - l?--(t) = A cos wIt + B sin wIt
C cos wit + D sin wit
A and C are obtained by evaluating the left-hand sides at t = O. Band D are found by evaluating the derivatives of the left-hand sides at t = O. utilizing (5.249). The result is
"++(') - e--(I) = ["++(0) - e--(O)J cos
•
and
dt 2 (l?+- - e-+) + wI (e+- - l?-+) = 0
l?+-(O - l?-+(O =
and
/,,/,'
e++(') - e--(') = ["++(0) - e--(O)] cos WI'
WI'
- i[e+_(O) - l?-+(O») sin
and
WI t
(5.258b)
(5.252)
e+-(I) - e-+(I) = ["+-(0) - e-+(O)J cos
WI'
- i[e++(O) - E?-- (0» sin
If we let the HI stay on for a time tl. we gel, then, for all later limes t that
WI t
. (M +(» t = I.,,,[ l!++( 0) - l!-_(O) ] Sill
Combining (5.252) with (5.247) and (5.248) gives
E?++(t):
I
and
1
'2 + '2 [e++(O) -
1
~[l?+-(O) -
Z
and
+~[£>++(O) -£>__ (0)] sin
wit
11ms following a 7r/2 pulse (5.254)
Suppose. now, we wish to calculate the expectation values of the x-. y-, and z-components of magnetization in the rotating frame. We utilize (5.159) (5.255') I ••
1"
= (M:(O» cos Witt WI I
(5.259a)
= Z[E?++(O) - l! __ (O)] cos WI tI
(5.253)
E?-+(O») sin wit
e-+(I) = 2["+-(0) + e-+(O)J + 2[e_+(O) - "+-(0)] cos
wlt l
= i(M:(O» sin Wltl
e __ (O)] cos wit
1
(5.258a)
(WI
(5.259b)
t I = "lf/2)
(M,(t» = 0
('»
(My = (M,(O» (M,('» = 0 •
(5.260)
which agrees with the classical picture of rotation about HI by rr/2 in the lefthanded sense.
'.9
Following a
"II"
pulse
(Wit)
="11")
(nle'(O)lm) = (nle
(M.('» = (My(l)) = 0 (M,(<) = -(M,(O» .
(5.261)
Again, the result agrees wilh Ihe classical result. Equation (5.257a) Iclts us useful facts about Ihe off-diagonal elemenls of
=0
unless
n=m=k
(5.265)
(kle'(O)lk) = (kle
ell: fH- must be nonzero for Ihere to be transverse magnetization.
a) b)
u+- must be pure real for the transvcrse magnetizations to lie along the
c)
u+- must be pure imaginary for the transversc magnetitation to lie along
d (mI' d' e (t)lm) = d'd (mlel m )
x-axis.
=
*~ !(ml"(O)I,:)(nll£jlm~
Ihe y-axis.
u-+ = Ae+ itJI (and hence u+- = Ae- itJI ) (M+(t» = l'flA(cos
q, + i sin tP)
SO that
(M.('» = 1hA cos ~
Thus ¢ gives the phase angle of the
in the
-
0'"
!ii(")e'(O)1ti('~ .....
C
D
- l£j(<)e'(O)l£i(l') + l£i(<)l£W)e'(O)llm)d,'. (5.266) ,
'"
.....
"
B
'
F
(5.263)
(ml,' (0)1t; (t')l£i (')Im) = ~)mle' (0)1 m')(m' 11£;(,')1£; (<)1 m) Since m'
(5.264b)
x-v magnetization
G)' j !mll"(O)1ti(t')1ti('~
(5.262)
(5.264.)
~
+
~e .consider first the lenus A and B. Since m i= k (Ihe rransilion is between two dislmCI stales), A and B both vanish according to (5.265). To treal tenns such as C, we write
where A is real,
('» = 1M sin
B
A
Since use of an HI which lies along Ihe x-axis will produce components of magnetization perpendicular to Ihe x-axis (assuming thennal equilibrium al t = 0), it will necessarily produce a purely imaginary u+- no matter what angle, Wltl, is chosen. To produce magnetization along the x-axis, HI mllst have a component nonnal to the x-axis. Such a magnetic field will produce a u+- with a nonvanishing real component. In general, if e-+ is given by
(My
- !mll£iln)(nle'(O)lm)!
x-v plane.
(mle'(O)lm') = 0
(m" k)
•
Ihe tenus C and F vanish, leaving
Jo
d 1 ' d' (mlelm) = h' [(mll£;(t')lk)(kll£;(<)lm)
+ (mll£;('JIkXk/l£j(,')lm)jdt'
S.lO Density Matrix - An Introductory Example
(5.267)
We convert Ihe malrix elements in the integrand by (5.107); Since the fonnal equations of the density matrix get rather involved, it is a good idea al this point to consider an example that will make ~he notation more concrete. We shall calculate the probability per second of transitions from a slate k to a state m. We consider thaI only state J.: is occupied at t = O. This assumption is not necessary, but it has the advantage, Ihat, with it, (dldtXmlulm) is directly Ihe probability per second of a lransition. Thus inilially all en's are zero except Ct;. As a result, at t = 0, the only nonvanishing element of the density matrix is (klulk), which is equal to I. By using (5.103), we see that 190
(mll£; (')1 n) = c(;!')(Em - E. )'(mll£ I (<)In)
.
(5.268)
Wc shall also adopt a convenient abbreviation. We shall define the quantum nU~bers m, n, and so on, in tenns of the corresponding energies, measured in radians per second: Em
T
sm
Eft
T
sk
~.
B.~
Thcn we have, using (5.268) and (5.269),
191
:'(mlelm)"
~2
t
J
The funclion Gmdr) is called the "correlation function" of "HI(t), since it tells how "HI at one time is correlated to its value at a later time. For a typical perturbation we have
(ml1i,(t')lk)(kl1i,(t)lm)e'(m-k)(t'-t)
o + (ml1i, (t)lk)(kl1i, (t') Im)e'(m-k)(t-t')Jdt'
(5.270) (5.277)
So far. our equations have been quile general as 10 the nature of the perturbation 1i 1(f). For example. it could vary sinusoidally in time. We shall assume for OUT example, however,that1f\(t) varies randomly in lime. By this we mean thai we shall consider a number of ensembles of systems of identical 110 and identical e(O). However, we shaU consider that1iI(t) varies from ensemble 10 ensemble. with properties described below. We shall therefore perfonn an average of the ensembles. We denote Ihis average by a bar. Then
+ (ml1i, (t) Ik)(k 11i, (t') Im)e'(m-k)( t-t')ldt'
(5.271)
As an example, suppose 'Ji\(t) were the dipolar coupling betwccn nuclear moments in a liquid. It might vary in lime. owing to the thennal motion of nuclei in the liquid. These motions would in general be different among various parts of the fluid, all al Ihe same temperature. We shall assume, moreover, that the ensemble average, such as
(ml1i, (t')lk)(kl1i, (t)lm)
depends on t and
Thus, if "HI (t) and "HI (t + r) were unrelated, we could average the two teons in the product separately, gelling (5.278)
=- 0,
However, we have, for r
(5.279) For typical physical systems, the penurbation "HI (t) varies in time, owing to some physical movement. For times less than some critical lime rc;, called the "correlation time", the motion may be considered negligible, so that hl(t) ';::' "HI(t +1'). For r > Te , the values of "Hl(t+r) become progressively less correlated (Q "HI(t) as l' is lengthened, so that Gkm goes to zero. Thus Gmk(r) has a maximum at r = 0, and falls off for 11'1 > TC;, as in Fig. 5.7. A function "HI(t) with the above propenies will be called a "stationary random function of time".
(5.272)
e only through their difference 1', defined by
Fig.S.7. Function G... t(T) (or lypical physie,,1 s)'slem
(5.273)
t-t'=T
II
T
That is, we assume (5.274) is independent of t but is a function both of T and of the pair of levels, m and k, with which we are concerned. The fact that (5.274) is independent of t is expressed by the statement that the perturbation is stationary. (A more complex case could be handled, but it would make everything much more complicated.) TIle dependence on 1', m, and k leads us to define a quantity GmkCr) by the equation (5.275) Since 11.1 (t) is a stationary penurbarion, we have that Gmk(r) = (ml"H1 (t)lk)(kl"H1 (t + T)lm)
" (k11i,(t + T)lm)(ml1i, (t)lk)
=- Gkm(-T) 192
(5.276)
Bearing in mind these properties, we now rewrite (5.271) in terms of Gmk(r):
d dt(mlelm) = -;.
I
J
[Gmk(r)e-i(m-k)T + Gmk(_r)ei(m-k)Tjdr
h 0 t
".!...2 JG mk ()e-l(m-k)'d T T h
-,
(5.280)
This equation tells us the rate of change of (mlelm) Te, the limits of integration may be taken as ± =, so thaI (d/dt)(mlelm) becomes independent of time. There is a range of 0 < t < Te; for which the transition rate is not constant. And, of course, if (mll?lm) becomes comparable to unity, we do not expect our pelUrbation approximalion to hold, since Ihe initial population (k1elk) would necessarily be strongly depleted. We shall now confine our anemion to times longer than Te;, assuming that for them (mlelm) has nOI grown unduly. Then we have 193
~ (mlelm) =
1 h
2
+=
J
GlIlk(T)e-i(m-k)r = Wkm
(5.281)
-=
where Wkm is the probability per second of a lransition from stale k to m. Equation (5.281) is closely related to the well-known resuh from timedependent perturbation theory:
W'm =
2: !(klVlm)I'e<Er)
l(ml'lt,(t)lk)I'· Gm,(O)
where (kIVlm) is the matrix element of the interaction between states k and m, and P(Er) is the density of final states in energy. In fact. by utilizing (5.275), we see that Wkm of (5.275) involves a product of twO matrix elements of the perturbation. In dlC present case the energy levels are sharp, but the perturbation is spread in frequency, whereas the usual lime-dependent perturbation is monochrommic but the energy levels are smeared. In view of the similarity to the usual time-dependent theory, we are not surprised that the tenns A and B of (5.266), which involve only one matrix elemen! of the penurbation, vanish. The integral of (5.281) is reminiscent of a Fourier transfonn. Let us then define a quantity Jmk(w) by the equation
+= Jmk(W)::
J
Gmk(T)e-il
(5.282)
-=
with the inverse relalion Gmk(T) ::
2~
kill
is independent of
TC'
1 Gmk(O) :: 2'1f
J
Jmk(w)eiwr dw
+=
J
h2
-=
which tells us that the area of the spectral density curve remains fixed as T c vanes. A sel of curves of Jmdw) for three different TC'S is shown in Fig. 5.9. A simple consequence of the fact that the area remains fixed as T c varies is found by considering Fig. 5.9. We note that if the frequency difference m - k were equal to the value WI. the spectral density Jmk(WI) of the curve of medium T c would be the greatest of the three at WI. Consequently, as T c is varied, the transition probability Wkm has a maximum. The maximum will occur when (m - k)Tc '::! I, since it is for this value of T c thai the spectrum extends up to (m ~ k) without extending so far as 10 be diminished. J m,
(5.283)
..-T,1ong
I
, J
J
\
o
(5.28la)
Now. in a typical case, the interaction matrix element (ml1t 1(Olk) runs over a set of values as time goes on. As one vanes something such as the temperature. the rate at which (ml1t\(t)!k) changes may speed upor slow down (Tc changes),
(5.284.)
Jmk(w)dw
-=
:: Jmk(m - k)
(5.284)
But. from (5.283)
+=
Jmk(W) may be thought of as the spectral density of the interaction matrix G mk . We expect that Jmk will therefore contain frequencies up to the order of IIT c (see Fig. 5.8). In tenns of Jmk' we have
W
but the set of values covered remains unchanged. As a physical example, the dipole-dipole coupling of a pair of nuclei depends on their relative positions. If the nuclei diffuse relative 10 each other. the coupling takes on different values. The possiple values are independent of the rate of diffusion, since they depend only on the radius vector from one nucleus to the other and on the spatial orientation of the moments. However. the duration of any magnitude of interaction will depend on the rate of diffusion. Then we have that
Wl
~
T"
short W
Fig.S.9. Jml(l
In the event thai Tc «::: I/(m - k). J",k(w) extends far above the frequency of the transilion. It is then often a good approximation to say that Jmk(m - k) '::! Jmk(O)
lei us define a quantity
"'it:. 5.8. Typical spectral density plOl
194
2(~)Jm'(O) =
0'
(5.285)
such that
+=
J Jm,(w)dw
(5.286)
-=
195
where 2(a/Tel is lhe width of a rectangular Jmk(W) whose height is Jmk(O) and whose area is the same as that of the real Jml:(w), From our remarks, a ~ I. Then we have, combining (5.284), (5.284a), and (5.286), T, h I(m I'H1(')lk)1 ' Jm.(O) = 2a
(5.287)
By utilizing (5.281a) and (5.285), then, we have
w
_ l(ml'HI(t)lk)I'
km -
h2
~ o'c
(5.288)
which holds for Tc < I/(m - k). Since one can often estimate the mean squared interaction as well as the correlation time, (5.288) provides a very simple method of computing the transition probability in the limit of shan correlation time (rapid motion). As a side remark, we note that for a flat spectrum Jmk(w). (5.288) will hold approximately for all Tc ~ 1/(m - k). We can use it to estimate roughly the maximwn rate of transition Ihat'H\(t) is capable of producing under the most favorable circumstances. by noting that this occurs when Tc ';;t I/(m - k). We get the approximate relation, then.
~ l(ml'HI(t)lk)!'
W)
km max -
"1
~
(k) am.
h~e-H/Tb
+ T)
(5.294)
where TO is lime defined by (5.295) in teons of the probability per second, that Hq(T) willjurnp from +h q to -It'!: I
W,
(5.295)
'lI
We assume for our example that this time is the same for all three components of the field. Oearly TO may be taken as the correlation time. 3 Substituting this form into (5.292) we find
., , '! (k Ilq Im) !" h q 1 2T," Jm.l.:(w) = "(nh +w '0
(5.296)
'"" 2 2
2
2TO 1l+(m-k) 2 2 TO
(5.297)
It is interesting to apply this formula to compute the T} for the case of spin For this case, as we saw in Chapter I, (5.291)
q,q'
where we have recognized that it is Hq(t) that varies from one member of the ensemble to another. For simplicity we shall suppose that the x-. y., and z-components of field fluctuate independendy. This means in effect that knowledge of Hz at some time is not sufficient to predict H y at that time. With this assumption. we get in (1.291) only terms for which q = q. Let us define +~
"H',("')"H'."(''''+'"'T'')e·-ilolTdT
(5.293)
The transition probability Wkm is then
+ T)lm)
=,~r.' L(mII,lk)(k1I,. Im)H,(')H,.(' + T)
196
k)
W.l.: m = [L. I'n hq l(mI 1q lk)1
Gm.(T) = (m!'H,(t)lkXkl'H,(t
J
Hq(t)Hq{t
(5.290)
The"
j~k(w) = l'~h21(mllqlk)12
;2 I:i:a,.,(m ,
Evaluation of j~k(w) requires information on the physical basis of the field fluctuation. Moreover, even when one possesses thal information, the malhemat~ ical problem of finding the correlation function may still be too difficult to solve. One can then often assume a function that, on general physical grounds, has about the correct behavior. For certain simple cases one can aClUally compute the correlation function. One such case arises when the field Hq(r) lakes on either of two values and makes transitions from one value to the other at a rate that is independent of the time since the preceding transitioo. Assuming the field takes on values of ± h q , we show in Appendix C that
(5.289)
HI (t) = -"(nh[H::t:(t)l::t: + Hy(t)Iy + H z(t)IzJ
-'or. q=r,y,z L H,(.)I,
W km =
-=2W
Our remarks so far apply to many forms of interaction 'H1(t). In order to be more coocrete, we now specialize. to consider a panicular fonn for the interaction 'HI (t). We shall assume that it consists of a fluctuating magnetic field with %-, y-. and z-eomponents that couple to the components of the nuclear moment:
=
Then we have
(5.292)
,
i.
(5.298) = 2Wt /, -1/' T, ' Assuming a strong static field in the z-direction, the matrix elements between the levels are -
I(W.I- ~)I' =
t
I(W,I- ~)I' = i I(~II,I
(5.299)
_ ~)I' = 0
3 Note thllt although we have not gi",:n II Pl'"ecise definition or the correlation time. once the correlation runction is specified, there is always a P'"ecisely definf!d parameter (in this case TO) thal enlers the problem and ehllfllCtC':n:res the time scale.
197
5.11 Bloch-Wangsness-Rcdfield Theory
We shall assume h z = h y = hz, so that
h; = !hij where hij =
hi + h~ + h;. Then, since m -
...!- = 2')'2 hij Tl
(5.300)
n 3 1
k = wo' the Larmor frequency (5.301)
TO
+wfiT6
This function is ploned in Fig. 5.10. We see that it indeed has the sort of variation we had predicted earlier from our general arguments based on the "constant area" under the curve of Jmk(W). We note thal the minimum comes when WOTO = 1.
We tum now to a more general treatment of the density matrix, following the ideas of Redfield [5.10], which are closely related to a treatment of relaxation due to Wangsness and Bloch [5.11]. All of the basic ideas are anticipated physically in the basic work of Bloembergen et al. [5.12]. The development of the basic equation of Redfield's theory is a generalization of the treatment given in the previous section for computing transition probabilities. Redfield shows that the elements of the density matrix obey a set of linear differential equations of the following fonn: 4
~I!:a' =
L
Ra(X',{3{3,ei(a-a'-{3+{3')tl!~{J'
(5.303)
{3,{3'
In T,
where R aa , ,{3{J' is constant in time. In this equation the time-dependent exponential has the property of making any tenns unimportant unless 0' - 0/ = {J - {JI. Therefore we could write (5.303) as Fig.5.10. Variation of TI with the correlation time "0
By utilizing this fact, we can calculate the minimum value of T 1 from (5.301) and obtain, in agreement with (5.289)
_I) __I ')'~h~
(5.302)
T} min - 3 WO
Now, when TO gets very long, the field Hz simply gives rise to a static line broadening (in this case, since only two discrete values can occur, the stalic line consists of a pair of lines at ± h z ) and is simply related to Ihis static width. Of course, wo is known for a typical experiment. Equation (5.302) therefore tells us the remarkable fact that knowledge of the rigid lattice line breadth, plus the resonant frequency, enables us to compute the most effective relaxation possible from allowing the line broadening interaction to fluctuate. In general the correlation time is changed by changing the temperature of the sample. Although (5.302) enables us to predict the minimum value of Tl' it cannot tell us at what temperature the minimum will occur unless, of course, we know the dependence of TO on temperature. We note in addition that the measurement of T} as a function of temperature can provide infonnation about the temperature variation of whatever physical process produces the fluctuations.
"&
~e:at = L'Raa',{3{3te~f1'
where the prime on the summation indicates that we keep only those tenns for whieh 0' - 0" = {J - {JI. The diagonal pan of this equation (that is, the part we should have if we kept 0' = ai, {J = {JI only) has the same fonn as the "master" equation, (5.13). The conditions under which (5.303) or (5.304) hold are given in tenns of R OtOt '(j{3" T c , and a time interval .1t. .1t defines a sort of "coarse graining". That is, we shall assume that we Ilever try to follow the details of e· for time intervals less than .1t. We must be able to choose such a time, subject to the simultaneous conditions that
.1t:» T c and
(5.305)
1 -;;--'---:» LIt
(5.306)
R aa ',{3{3J
Equation (5.305) will permit us to set the limits of certain integrals as ± 00, as we did in (5.281). Equation (5.306) will guarantee that, during .1t, the density matrix does not change too drastically, so thai our perturbation expansion has validity. Since the R aa ',{3{3"s are comparable to the inverse of the relaxation limes, these conditions are equivalent to saying that Tj and T2 are much longer than T e . These are also the conditions of motional narrowing. The reader will pehaps note that these are the very conditions stated in Section 2.11, which must hold if simple time-independent transition probabilities are to hold. What we are doing here is to generalize Ihe usual time-dependent perturbation theory so that it includes the coherence effects associated with phase factors in the wave function. 4 For compactllCS5 in writing, we use the notation eo,,"
(aiel a'). 198
(5.304)
for the matrix clement
199
The beauty of (5.304) is that it provides us with a simple set of linear differential equations among the elements of the density matrix, which in principle we can always solve. They wiJIlead to a set of "nonnal modes". We note that there is a great deal of similarity to the rate equation describing population changes. Moreover, expressions for the Raot'(J(J"s are given by Redfield's theory, so that we have the relaxation times given in tenns of the atomic properties. Before outlining the derivation of (5.303) and (5.304), we remark that there are twO ways of utilizing (5.303) or (5.304). In the first method, one solves for the behavior of each separate element of the density matrix and then computes the time dependence of the physical variables of interest (such as the x-components of magnetization M z ) by the fundamental equation
(Mz ) =
L: e••,(a'IMzla)
(5.307)
a,o' The second method involves seeking a differential equation for (Mz ). One does this by writing
d-d '" ' dt (Mz) = dt L.J {laQ',(a IMzla) a,a' =
.,.L:
,
d:" = *[,(0), :>lj(t)] +
G)' Ja
[[,(0), :>lj(t')], :>lj(')]dt'
d'
(5.308)
We then use (5.303) to express the time derivative (dldt){laQ'" In fact, since (5.309) we have
+ ei(a-a')! d{l(.ta' dt
(5.313)
We compute the 0'0'1 matrix element. There will be contributions from bOlh tenns on the right. We consider first the contribution of [{I·(O),1ij(t)]:
(alle'(O), :>li(t)lIa') = Dale'(O)IP)(PI:>lj(tl!a') p
- (al:>lj(t)IPJ(Plg'(O)la')
(dg·"')(a'IMz1a)
d{l~a' = i[a _ 0"].
these equations are fewer in number than are the equations of the elements of the density matrix, their solution may be considerably simpler than the original set. This trick will work when the relaxation mechanism and operators are such thai the expectation values of the operators pick out only a small number of the possible nonnal modes. We shall illustrate this use of (5.312) shortly. First, however, we tum to a description of the derivation of Redfield's fundamental equation. Our starring point is the basic equation for the time derivative of fl·, (5.111):
(5.314)
We shall now introduce the idea of an ensemble of ensembles whose density matrix coincides at t = 0 but whose penurbations 1i1(t) are different. (We are therefore flot allowing an applied alternating field to be present. That is, we are describing relaxation in the absence of an alternating field. One could add the effect of the alternating field readily. We discuss it in the next section.) We shall assume that the ensemble average of 1i 1(t) vanishes. This amounts to our assuming that1iI(t) does not produce an average frequency shift.~ Let us discuss this point. In general we expect
(5.310)
(5.315)
This relation enables us to transfoml (5.303). By substituting into (5.303) and utilizing (5.309), we find
where J(q is a function of the spin coordinates, and Hq(t) is independent of spin. For example, we saw that hi (t) had this fonn if it consisted of the coupling of a fluctuating magnetic field with the X-, y-, or z-components of spin. In that case q took on three values corresponding to the three components x, y, and z. If hI (t) represented a dipole-dipole interaction of two spins. there would be six values of q corresponding to the tenns, A, B, ... , F into which we broke the dipolar coupling in Chapter 3. Since we are dealing with stationary penurbations, the ensemble average of 1ij(t) is equivalent to a time average. In general we assume that the time average of Hq(t) vanishes, causing 1ii(t) to have a vanishing ensemble average. As a consequence we shall set
dt
d{lQ'a'
d'
fl aa'
= i(a' - a){laa'
+ L RaQ"
,{3{3' {I{3(J'
P,P'
,
= -,. [I?, 7-lo]Q'Q" +
L
Roo',pp'I?PP'
(5.311)
P,P'
l
We then substitute this expression into (5.308), obtaining
d(Mz ) d'
---=
L:
a,Q",pjJ'
(*[e, :>lo] ••' + n••"pp,gpw)(a'IMz1a)
(5.312)
Although it is not obvious from (5.312), under some circumstances the righthand side is proponional to a linear combination of (Mz(t», (My(t», and (AdAt», giving us a set of differential equations similar to those of Bloch. If 200
(5.316) ~ If there is zero shifl.
a
shifl,
we
<;ao
include
the lwerage shifl in ho, redefining 1-£1 (t) lo give a
201
where the bar indicates an ensemble average. This means, we have remarked, that we cannot let 1-£1 (t) be a time-dependent driving field such as that applied to observe a resonance. On the basis of (5.316) the first teon on the right of (5.313) vanishes when averaged over an ensemble. We proceed in a similar way to compute the aa' matrix element of the second teon on the right of (5.313). By utilizing the fact that
=~"...."...,. . . . Hq(t)Hq,(t + T)e-i... r dT
J
Lqq,(w):=
(5.324)
o By utilizing the fact that Hq(t)Hq,(t + r) is real' and is an even function of T, it is convenient 10 define the real and imaginary parts Lqq,(w):
Re {Lqq,(w)} =
,+=
2"
J·H·,"(t~)H~,-,(~t-+-T')
cos (wT)dT := kqq'
-=
(5.317)
(5.325)
J
=~~~~~ 1m {Lqq,(w)} = Hq(t)Hq,{t + T) sin (wT)dT
and defining T=t-t'
o
(5.318)
Since the only imponant contributions to (5.319) come from teons that satisfy the condition a - a' = fJ - If, we can combine the first twO teons on the right of (5.319) as
-!, (L: L: {(all<'I,8)CP'I[(" la')lk,.,(a - ,8) + k,~(a' h
/l')J
fJ,flq,q'
x e'(a-ft+p'-<>,), e~w
j)
(5.326)
The last two teons of (5.319) are
-!, L:
h fJ.fJ',q,q'
{e~ft 1
+ep..,I
(5.321)
is independent of t and goes to zero when T exceeds some critical value T e. In this case we can consider times t grealer than T c, peonitting us to extend the upper limits of integration to T = +00. We now define the correlation function G(l/(3(l/'(3'(T) as G,pa'W(T) "
(al?i, (t)IP)(P'I?i, (t +T)la')
(5.322)
GaPa'W(T) = L:(all<'IP)
We then define the spectral density Lqq,(w) of the interaction as
202
(5.327)
The imaginary pan of Lqq' can be shown to give rise to a frequency shift corresponding to the second-order frequency shirt of a staLic interaction. We shall neglect this effect, keeping only teons proportional to Re {L qq ,} because they give the relaxation. Therefore we replace the Lqq,'s by the kqq,'s. In analogy to our earlier discussion we now define the spectral densities J(l/(l/'{1IP(w) as (5.328)
-= Then, combining (5.326), (5.327), and (5.328), we have de:o:' _ "
~
By utilizing (5.315) we have that
/l'»)ei(a-P')'j .
- LJ
R
O'o-'IJ(3'
e i(o--O"-/H(3')L".
<:/J{1'
(0)
(5.329)
(3,{1'
(5.323)
where R(l/O",/J/3' is given as 6 As long as the 1<'s arc taken as Ilermitil\ll 0l)erl\tors, the U,'s arc real. If one ehooses the 1<, 's lIS non-Ilerrnitian, lhe /f, 's become complex, but the only f'ft' 's that are nonzero then involve q and q"s which make U,(t)II,,(L + r) real.
203
(5.334) (5.330) Equation (5.329) relates de"ldt at time t > T( to eO al t ::: O. II is the first leon in a power-series expansion. In order for the convergence to be good. it must imply that f!ppl at time t nOI be vastly.different from this value al t = ~. This implies thai we can find a range of tImes t such that t:> Tel but sull 1!~1J,(t) ':l!' Upp,(O). This laner condition implies that 1
~--'---
R".. ,~W
:» ,
(5.331)
The important trick now is 10 nOle thai if (5.331) holds true, we can replace l'flIJ'(O) by i!P13,(t) on the right side of (5.329). By this step, we convert (5.329) into a differential equation for [/, which will enable us 10 find [/ by "integration" at times so much later than t <:: 0 that l?PIJ,(t) will no longer be nearly its value at t = O. The resulting equation is (5.303). The physical significance of our conditions is now seen to be that we never ask for infonnation over time intervals comparable 10 Te, and that in this time interval the density matrix must not change too much. In practice. this implies that
TI. T2
>Tc
(5.332)
As we shall see in greater detail, the condition Tc < T2 is just that for which the resonance lines are "narrowed" by the "mOlion" that produces the fluctuations in 'HI (t). Because ~oJJfJ = R/J/J,OOl
(5.333)
(that is, the transition probability from 0' to fJ is equal to that from fJ to 0'), the solution of the Redfield equation is an equal distribution among all states. This situation corresponds to an infinite temperature. Clearly the equalions do not describe the approach to an equilibrium at a finite temperature. The reason is immediately apparent - our equation involves the spin variables only, making no mention of a thennal bath. The bath coordinates are needed to enable the spins to "know" the temperature. A rigorous method of correcting for the bath is to consider that the density matrix of (5.313) is for the total system of bath and spins. Since in the absence of HI the spins and lattice are decoupled, we may take the density matrix to consist of a product of that for the spins, a, and that for the lattice, ~L. We take for our basic Hamiltonian '110 the sum of the lattice and the spin Hamiltonians (which, of course, commute). 'HI commutes with neither and induces simultaneous transitions in the lattice and the spin system. Then we have 204
Introducing spin quantum numbers sand !l, and lanice quantum numbers / and /', we replace 0' by sf, and so on. Then we assume that the lattice remains in thermal equilibrium despite the spin relaxation:
ll}1' = Off'
e-"/IkT "/"/kT
Ee
(5.335)
I" We then find the differential equation for
d
L• •
I..d.
r
d/ ll /I'uu ') = o/f'llfl' dtan'
(5.336)
and sum over f. The result, in the high temperature limit, is simply to give a modified version of Redfield's equation, with density matrix for spin a replaced by the difference between u and its value for thennal equilibrium at the lattice temperature q(T). We therefore simply assert that for an intcraction in which the lattice couples to the spins via an intcraction 'HI (which, to the spins, is time dependent),7 the role of the lanice is to modify the Redfield equation to be
du;o-' ~ +i(Ol-Ol'-.B+.8')t.. - d - = L. Roo-',fJ.B,e [ufJ.B' - u.BfJ,(T)]
t
(5.337)
.oft'
where a,a',fJ,' stand for spin quantum numbers, and where u.8.8,(T) is the thennal equilibrium value of ufJ.8':
C"/J/ kT
"~~,(T) = 6~W Le WIlT
(5.338)
~"
That (5.337) should hold true is not surprising in view of our remarks in Chapter I concerning the approach to thennal equilibrium. We note here. however, that our remarks apply not only to the level populalions (the diagonal elements of u) but also to the off-diagonal elements.
7 "HI involves both spin Rnd lattice coordinRlcs. If we treat the lattice quantum mechanically. the lattice variables are operators, and 'HI dOClS not involve the lillle explicitly. If we treal the lattice classically, 'H\ ;n\'oln'$ the time c:o:plicity. Thill this must be so is evidcnt, since the coupling must be time dependent to induce spin transitions between spin stales of different energy. However, it is time indcll·cndent when the lattice ml'lkcs a simultaneous transitiOIl that just absorbs the spin energy. 205
5.12 Example of Redfield Theory
where q = x, y, z, and
We tum now to an example to illustrate both the method of Redfield and some simple physical consequences. The example we choose is Ihal of an ensemble of spins which do nOI couple to one another but which couple to an external fluctuating field, different al each spin. The external field possesses X-, y-, and z-componenls. This example possesses many of the features of a system of spins with dipolar coupling. However, it is subslanlially simpler to lTeal; moreover, it can be solved exactly in the limit of very short correlation time. For the case of dipolar coupling, Ihe nUClUations of the dipole field arise from bodily motion of the nuclei, as when self-diffusion occurs. The correlation time corresponds 10 the mean time a given pair of nuclei are near each other before diffusing away_ OUT simple model gives the main qualitative features of the dipolar coupling if we simply consider the correlation time 10 correspond to that for diffusion. In panicular, then. our model will exhibit the imponant phenomenon of motional narrowing, which has been so beautifully explained in the original work of BJoembug~n et al. Before plunging into the analysis. we can remark on cenain simple features that will emerge. At the end of this section we develop these simple arguments funher. showing how to use them for more quantitative results. We may distinguish between the effects of the Z·, y-. and z-eomponents of the fluctuating field. A component Hz causes the precession rate to be faster or slower. It. so to speak. causes a spread in precessions. It will evidently nor contribute to the spin-Ianice relaxation because that requires changes in the com· ponent of magnetization parallel to Ho. but it will contribute to the decay of the transverse magnetization even if the fluctuations are so slow as to be effectively static. In fact, as we shall see, it is Hz that contributes to the rigid-lattice Line breadth. The phenomenon of motional narrowing corresponds to a son of averaging out of the Hz effect when the fluctuations become sufficiently rapid. The z- and y-components of fluctuating field are most simply viewed from the reference frame rotating with the precession. Components fluctuating at the precession frequency in the laboratory frame can proouce quasi-static components in the rotating frame perpendicular to the static field. They can cause changes in componenls of magnetization, either parallel or perpendicular to the static field. The former is a T 1 process; the latter, a T2 process. Clearly the two processes are intimately related, since the magnetization vector of an individual spin is of fixed length. The transverse componentS of fluctuating magnetic fields will be most effective when their Fourier spectmm is rich at the Larmor frequency. For either very slow or very rapid motion, the spectral density at the Lannor frequency is low, but for motions whose correlations time, is of order I/wo, the density is at a maximum. The contribution of H~ and H y to the longitudinal and transverse relaxation rates therefore has a maximum as , is changed. Let us consider, then, an interaction 1t1(t) given by
'H\(t) =
-,"h L:Hq(t)Iq
(5.339)
1to = -"(n/tHol: = -liwolz
(5.340)
where. wo is the Larmor frequency. We characterize the eigenstates by a, the eigenvalues of (5.340). These are wo times the usual m-values of the operator I z . (Herem = I, I-I, ... -I). We shall continue to use thenotalion a,d,fJ,fJ', however, rather than m, in order to keep the equations similar to those we have just developed. The matrix elements (al1tI(t)la') are
(al'H,(')la') =
-,"h L:Hq(')(allqla')
.
(5.341)
q
Then the specrral density functions JOtpo'P'(w) are
We now use the symbol kqq.(w) introduced in the preceding section as +~
kq¢(w) =
4J 'H'q"(")H"q-,T.('--:+-:T">e- iw.. d,
.
(5.343)
-~
Clearly the fluctuation effects, correlation time, and so on are all associated with the kqq"s. For simplicity let us assume that the fluctuations of the three components of field are independent. That is, we shall assume
Hq(t)Hq,(t + r) = 0
if q "" q'
(5.344)
For example, (5.344) will hold true if, for any value of the component H q , the values of H q• occur with equal probability as IRq'1 or -IHq,l. We note that kqq(w) gives the spectral density at frequency w of the q-component of the fluctuating field. With the assumption of (5.344) we have, then, (5.345) We now seek to find the effeci of relaxation on the X-, y-, and z-components of the spins. To do this, we utilize the second technique described in the preceding section, that of finding a differential equation for the expectation value of the spin components. Let us therefore ask for (dldt)(Ir ), r = x,y, or z. By using (5.312), we find
d(I,) - = " L. i h[ ll,1to Joo,,(a' I.. 0'l l+) " L. Roa',f3fJ'llf3J3'(U ' IIrla) dt 0,0' o,o',P,P' (5.346)
q
206
207
"
+1'~
The first tenn on the right, involving 'Ha, can be handled readily:
L , 2.r.Q. 'H'o)Cf(f,(a'IIr 10) h
= _,i Tr {(g1iO
,
0,.
-1toe)Ir}
=
Ii Tr {eHoI, -
=
Ii Tr {e[Ho,!,]}
e I , Ho}
= -i'YnlIo Tr {e[I~, I r ]}
(5.347)
If r = z, this term vanishes. If l' = x, we have
-i1'nHO Tr {ily!?}
= +1'n Ho(Iy )
(5.348)
If r = y, we get --YnHo(l:t'). Thus we have
L 0',0"
*[e, 'h'o]O'O',(c/llrlo)
= 1'n«1) x HO)r
(5.349)
RO'O",(3(3'!?(3/3,(O"IIr la)
.
(5.350)
0',0" ,/3,(3' As we have seen, RCl'O" ,(3(3' is itself the sum of four tenns [see (5.330)]. We shall discuss the first teon. JO'(3O"(3'(O" ~ 13'). By using (5.345), we find
L
~
1; I: L
RO'O', ,(3(3' e(3(3' (ollIr 10)
0',0" ,(3,/3'
1; I: (PIIqlo)(oII,Iq -
= i1'~ku(wo)Tr {I:t'Iye - I:t'ely} =
i-y;ku(wo)Tr {U:t'I y - IyI:t')e}
= --y;k:t':t'(wo)Tr{IzU} = -1'~k:t':t'(wo)(lz)
(5.352a)
The tenn q = y gives, in a similar manner,
-1';kyy (wo)(Iz)
.
(5.353)
RO'O', '/3/3'!?/3/3'('/ 1I z 10) = --y;( k:t':t'(wo) + kyy(wo) J(I z )
. (5.354)
By combining (5.346), (5.349), and (5.354), we have
({3'jlq]O")(o'jlrlq!?I,O')kqq(o' - {1)
(5.351)
where the last step follows from the basic properties of orthogonality and completeness of the wave functions 10), and so on. We are able to "collapse" the indices 0 and .0. but we cannot do the same trick for 0' and p' because they occur not only in the matrix elements but also in the kqq's. In a similar way one can obtain expressions for the other three [enns in RO!O!',(3(3" getting finally
=
DW, 10)(011[/" I,lellPlk,,(o - P) ',p = ,,~[ I:
L
O",(J' ,q
L:
,~
0',0" ,(3,(3'
(oIIqIP)(P'IIqlo')(P1dP')(o'II,lo)kqq(o' - P')
O',O",/3,(3',q
"" -y;
where, in the last step, we have utilized the faci that kqq(w) is an even function of w. To proceed further, we must now specify whether r is x, y, or z. First let us consider r = z. Then, since lr will cornmUie with [z. we get nothing from q = z in the last line of (5.352). Since matrix elements of ['" vanish except for L1m = ± I, the only states (0' and (3) that are joined by I q for q = x have 1,6 - al = wo, the Larmor frequency. Since [Iz.I z ] = iIy and [Iz,I y] = ilz , we have, then,
All told, then,
JO'(3O"/3'(o/ - ,O')e(3(3' (0"1 Ir 10')
2h 0',0",(3,/3' =
(5.352)
0)
',p
which is the driving tenn of [he Bloch equations describing the torque due to the external field. The second tenn on the right of (5.346) involves the relaxation terms;
L
I: (Wqlo)(ol[[I" IqJ, eIIP)kqq(P -
=.,~
j
=
Irlq}IP)kqq(fJ - a)
O',fJ,q
j
-i1'nHO Tr felIz, I:t'J}
:L ({3IIq]a)(aleUqlr cr,fJ,q
IqI,)eIPlkqq(o - P)
z d(I ) "" -y«1) x Ho): dt
-1'~[k:t':t'(wo)+ kyy(wo)](Iz)
(5.355)
This equation relaxes toward (Iz ) "" 0 rather than the thennal equilibrium value 10· To remedy the silUation, we should replace U by e- f!{T), as discussed in rhe preceding section. This result simply makes (Iz) relax toward the thennal equilibrium value 10:
d~,)
= 1((1)
x Flo), -
1~[k,,(wo) + kyy(wo)I«I,) -
10)
(5.356)
This is clearly one of the Bloch equations, with T 1 given by the expression (5.357)
O!,(3,q 208
209
We can proceed in a similar way to find the relaxation of the x-component. For it, the value q = x contributes nOlhing, but q = y or z does. The situation for q = y is similar to the one we have just discussed, connecting states 0' and {J which differ by woo On the other hand, when q = z, the states 0' and {J are the same (Iz is diagonal), so that 0' ~ (J = O. The spectral density of Hz at zero frequency enters. Therefore we find
L
R ua' ,PP'(JPfJ,(O"II:l: 10') = -'Y~[kyy(wo) + ku(O)](I:l:)
dMx
&
Mz
M; MJ
,
We now introduce the precessional motion by replacing M:l: and My (5.359)
For this equation there is no effect of replacing (J by (J - (J(T), since in thennal equilibrium (I:l:) = O. Equmion (5.359) and a similar one for (Iy ) look very much like the transverse Bloch equation describing a T2 process, except that the T2 for (I:l:) differs from that for (Iy ). Labeling these as T2:l: and T2y we get
1
= 'Y~[kyy(wo)+ k:z(O)]
(5.360a)
-1' = 'Y~[k:r::l:(wo) + kzz(O)]
(5.360b)
1',. 1
"
If we did not have the driving tenn from Ho, these equations would cause (I:l:) and (I y ) to decay with the rates T2z and T2y respectively. However, in general the precession rate wo is much faster than I/T2 :r: or I/T2y. Thus, we must average the T2 effects over the precessional motion. That is, the amplitude of the transverse magnetization will decay slowly compared to the precession frequency. h is straightforward to detennine this rate. We define the transverse magnetization M1.. by the equation
M1.. = iMx + jM y
(5.364)
= -T2x - T2y
NO):l: - 'Yn[kyy(wo) + ku(O)](I:l:)
so that
(5.361)
Theo dMl=2M dM1.. dt 1.. dt =2M dM:l:+2M dM y Zdt Ydt From (5.359) and (5.360), we get
(5.363)
dM1.. M2 M2 M ~ -dt = lvfxwoMy - _z - MywoM:l: - - " T2z T2y
(5.358)
which gives for the derivative of (I:l:):
-
My -woM:l: - T2y
Using (5.363) and (5.362b), we get
u,u',fJ,P'
d(l.) ----;u= 1'n«1) x
dM y
ili =
= woM:l: - T2x
(5.362.)
M x = M1.. cos (wot + 4»
(5.365)
My = -lvh sin (wot + 1') in (5.364), getting
lvh dM1.. = -Ml
(COS 2 (wot + r/J) + sin 2 (wot + r/J»)
~
~x
~y
(5.366)
Averaging this equation over one period, we replace cos 2 (wot+r/J) and sin 2 (wot+ r/J) by getting
!
dM1.. dt
=
_1_)
-M1.. ~(_I_ + 2 T2z T2y
= _
M1.. T2
-!... _ ~ (_1_ + _1_) T2 - 2 T2z
(5.367)
(5.368)
T2y
Combining (5.360) with (5.368), we see that 1 1 T2 = 2Tl + 'Y;kzz(O)
(5.369)
Our relaxation mechanism therefore leads to the Bloch equations. Of course one cannot expect that in general the Bloch equations follow from an arbitrary 7-tl(t), and one would have to study each case to see whether or not the Bloch equations resulted. To proceed further, we need to know something about the spectral densities of the x-, y-, and z-components of the fluctuating field. We shall once again assume a simple exponential correlation function, with the same correlation time 7"0, for q = x, y, and z: Hq(t)Hq(t +7") = H~exp(-I7"I/To)
(5.370)
which gives (5.362b)
-
kqq(w) = Hi
TO
I +W
2 2
(5.371)
7"0
In tenns of (5.371) we have, then, 210
211
'1
f
relative to its neighbors by diffusion. In Ihe time phase angle 6¢ over its nonnal precession:
6¢
=
T,
a spin will precess an extra (5.315)
±l'nlH.:IT
(5.372) After n such intervals, Ihe mean square dephasing tJ.¢2 will be
1 '(2
;;:;- '" '111 ..I2y
tJ.¢2 = n6¢2 = n1'~H;T2
T() ) BlTO + "2" Hz I +W2T2
The number of intervals n in a lime t is simply
0 0
from which we gel
t
'n
(5.311)
n=-
1 1 =..,2H2 ro + T2 In l 2T) ;;; ....2
(5.316)
T
(H2z TO + ~(H2 + 82) TO ) 2 z Yl+w2r2 o
(5.373)
If we take as T2 the time for a group of spins in phase al t = 0 radian out of step, we find
0
10
get about one
(5.318)
We nOle first of all that Tl goes through a minimum as a function of TO when woTQ '"
1. The nuclualing fields Ihal determine T 1 are the z- and y-components
at the Larmor frequency. If we view the problem from the rotating frame (that is, one rotating at the Larmor frequency), these results are reasonable. since the T, corresponds to a change in the z·magnetization. Such a change is brought about
by "static" fields in either the z- or y-directions in the rotating frame. since in the rotating frame the effective field is zero (HI. of course, is zero). But "static" x- or y-fields in the meating frame oscillate at WQ in the laboratory frame. On the orner hand, lhe decay of the x-magnetization must arise from the "static" y- or z-fields in the rOiating frame. Since the z-axes of the laboralory and rotaling frames coincide, for the z-field it is the static laboratory component that counts, but for the y-field, il is the laboratory component at the Lannor frequency Ihal is important. We note that in the limit of very rapid motion (wOrD <:: 1), and assuming an isotropic fluctuating field
H'• • H'y = H'•
(5.314)
Tl and T2 are equal. Physically, for our model, this result signifies that for a very short correlation time, the spectral density of the fluctuating field is "white" to frequencies far above the L1.rmor frequency, so that the x', yo, and z-directions in the rotating reference frame see equivalent fluctuating fields. The two terms in the expression for T z have simple physical meanings. One tenn depends on H~. It represenls the dephasing of the spins due to the spread in precession rates arising from the fact that H ~ can aid or oppose Ho. This term can be readily derived by a simple argument, which we give below. The second term, as we shall explain, results from broadening pf the energy levels due to the finite Iifelime a spin is in a given energy state. Let us now tum to a simple derivation of the first term of the equation for T2. We assume the field as a value lH~1 for a time r; then il jumps randomly to ± IH~I. Such a change in field in practice arises because a nucleus moves
'I'
...!... T2
= ",,2 H 2 T In
r
(5.379)
We note that the shorter T (Ihal is, the more rapid the mOlio~), the narrower Ihe resonance. This phenomenon is therefore called motional narrowing. We see that Ihe motion narrows Ihe resonance because it allows a given spin 10 sample many fields H~, some of which cause ilia advance in phase; olllers, 10 be retarded. The dephasing takes place, lIlen, by a random walk of small sleps, each one much less Ihan a radian. In conttast, when there is no motion, a given spin experiences a constant local field. II precesses either faster or slower than the average, and Ihe dephasing of a group of spins arises from Ihe inexorable accumulation of positive or negative phase. The contraSI with "collision broadening" of spectral lines is greal. In thai case Ihe phase of the oscillation is changed by each collision. Since the frequency is unperturbed belween collisions, Ihere is no loss in phase memory except during a collision. Since each collision gives a loss in phase memory, a more rapid collision rate produces a shorter phase memory and a broader line. With motional narrowing, there is no phase change when H r is changing from one value to another because the change is very rapid, but lIlere is a phase change during the time a given value of H r persists. More rapid motion diminishes the loss in phase memory in each inlerval. We have considered just one (enn of our expression for T2. The other term, clearly has the same dependence on TO as does the spinwhich involves lattice relaxation. We interprel it as the bro,1.dening of the line due to the finite life of a spin in any eigenstale as a result of the spin-Ialtice relaxation. The lifetime
H;,
8 Of course the word ~motioll~ here refers to tn.nstation of the position of the nudeWl, not to the change in spin orientation. 213
is finite because a field in the V-direction can change the .z-magnetization. We should estimate the order of magnilUde of the lifetime broadening to be
.cl.E =!!... T,
or
(5.380)
c,E I L!w=-=-
h T, Assuming isotropic fluctuating fields, we see that our example actually gives _I = _I + _ 1 (5.381) T2 T 2, 2T1 where T2, is the broadening due to the spread in the .z-field. The quantity 1fl'2', is often called the sUlllar broadening; the term lilT) is called the nonsecular or lifetime broadening. More generally, we replaced lilT) by IIT I" where T I, (which gives the nonsecular broadening) is related to T). As we have remarked, if we consider the secular broadening, we nOte that as TO decreases, T2 increases, oc the line narrows. Of course, we have seen that as one increases TO (slows the motion), the validity of the Redfield equations ceases when T2 ";:! TO. For longer TO'S, we cannot apply the Redfield equation. The longest TO for which the Redfield theory can apply, then, is TO = T2 , or , 82TfJ _1 = _1 = "'( TO T2 nz
or
7n!H,lrO = 1
..!....
(5.383)
So far we have excluded applied alternating fields from the time-dependent coupling 'H)(t). Let us now assume that such fields are present, giving an extra tenn, 'H2(t), in the Hamiltonian. We can include its effect in a straightforward way, as has been shown by Bloch. Introduction of 1t2(t) simply replaces 1tj(t) by Hj(t) + Hi(t) in (5.313):
0'
t 1" T',
Secular (1~) and non$OCular (Tn broadening versus TO. For
Fig. 5.11.
Toe!l/wo
I
where TOO is the value of TO for infinite temperature. The temperature variation of T t or T2 gives one a convenient measure of E and TOO. The narrowing smdies of Andrew and Eades [5.13) performed on molecular crystals, provide one such example. Another one is the work of Holcomb and Norberg [5.14) on self-diffusion in the alkali metals, and subsequently, Seymour [5.15] and SpoJcas (5.16) on aluminium. Here one has the interesting fact that, using resonance, these workers could measure the self-diffusion rate of both Li and AI (for which there is no radioactive isotope for use in the conventional tracer technique).
5.13 Effect of Applied Alternating Fields
T',
T',
(5.384)
(5. 382)
As we can see from our simple model, Ihis is just the value of TO at which a Iypical spin gets one radian out of phase before there is a jump. For longer TO'S, the spins are dephased before there is a chance for a jump. That is, they do not dephase by a random walk. The line breadth is then independent of the jump rate, giving one the temperature-independent, rigid-lattice line breadth. The twO contributions to the line breadth (secular and nonsecular) are plotted in Fig. 5.11. If one analyzes the relaxation via other mechanisms, the same general features are found. The fact that more than one transition may be induced will make important the spectral densities at frequencies other than 0 and wOo Often, 2w0 /11
comes in. For example, when the relaxation arises from the coupling of one nucleus with another by means of their magnetic dipole moments, the E and F terms of Chapter 3, which involve the product of two raising or IWO lowering operators, connect stales differing in energy by 2hwo. Our formulas show us that the measurement of T 1 and T2 will enable us to detennine TO. When the fluctuations in the interaction 'Ht(t) arise from bodily motion that varies with temperature, we can use resonance to study the temperature variation of TO. Often one has a "barrier" to motion and an activation energy E such that
the eXilmple in the led, T[
=2Tl
d'
.
:, = *le'(O),Hj(t) + H;
+
2'
(*) Jo 11.(0), (Hj(t')
+ H;
The effect of Hi in the first tenn on the right vanished when averaged over an ensemble. For that reason we were forced to consider the second tem. In general the contribution of 'H 2(t) to the first term does 1I0t vanish, since 'H 2(t) is identical for all members of the ensemble. If 1t2(t) is not too strong, we expect Ihat the first-order tenn is all that is necessary, and therefore we neglect the role of 'H 2 in the integral. Physically, this approximation amounts to our saying thllt
dt/
de·)
dt = dt 1f.
de')
1
+ ill n'!lax
(5.386)
where de*/dt)1f.2 is the rate of change of e· due solely to 'H2, and de*/dt)relax is
Toe!I/'r"IH.1 214
215
the rale of change of r/ we should have if 1t2 were zero. What we are neglecting are, therefore, nonlinear effects in the interactions.
Under what circumstances do we expect this approximation to work? The answer is that neither perturbation must change
'l
t (which is the upper limit of lhe integral), for the
'Hi
too much during the time
'Hi
[enns in the integral
is acting on a (/. which is not in fact e-(O) bUI must be corrected for the driving by 'Hi. Since we wish to choose t as about Te. this represent the fact that
requirement means that
1("I?t,I,,')IT<
<0:'
(5.387)
"
If 1t2 is too large to salisfy (5.387), we should then attempt 10 solve first for the combined effect of 'Ho and 'H2. using perturbation theory for 'HI- For example,
we may use a rotating coordinate transfonnation rather than a transfonnation 10 the intemcrion representation, thereby converting 11i(t) to a static interaction, which could then be removed by a second transformation to the interaction representation of the effective field. It is interesting to note that there is in fact a very close similarity between the interaction representation and the usual rotating coordinate transformation that renders H I a static field. Both are transformations to rotating coordinate systems. The interaction representation is a transformation to a system rotating at the Lannor frequency, whereas the usual transformation goes to a coordinate system rotating with H t . For simplicity we shall assume that (5.387) is satisfied. Note that since T, and T2 are much longer than T c for all our equations to be valid, (5.387) can still be easily satisfied evcn under conditions of saturation. There is one further consequel'lCC of the addition of a tenn 112(t). We have remarked that when we treat the latticc classically, the density matrix relaxes to its value at infinite temperature rather than to the thennal equilibrium value
e!:T)o f!{T) =
(5.388)
Z(T)
where Z(T) is the sum of states. When the change of 1i2 is small during the time T e , which characterizes the "lattice" motion, we expect that 1i2 looks like a "static" coupling to the lanice and that we should assume that the system relaxes at each instant toward the instantaneous density matrix:
f!{T. t) = Z(T. t) exp (-(?to + ?t,(t»)/kT)
.
(5.389)
This equation can be verified by treating the lattice quantum mechanically. When T e is long compared with the period of 1i2' (5.388) applies. Under circumstances where the Bloch equations hold, the shon T e leads oflen to T l = T2 , and the Bloch equations become
216
M x H() t+ Mo-M Tj
where
(5.390)
Mo = xoH(t)
(5.391)
and H(t) is Ihe instantaneous applied field. Explicit solution shows that the solution of (5.390) differs significantly from the nonnal Bloch equations (assum+ ing T 1 = T2, but Mo = xoMo) only when the line width is comparable to the resonance frequency. We can combine (5.386) with (5.311) to obtain the complete differential equation for the density matrix, including an applied alternating field: dUoot' i d - = -,,(Eo-' - Erx)l!rxtY t
' + -,i '" L.. (l!rxtY'(o "I 112(t) I0) l 0"
- (,,(?t,(')I,,")eo""'] +
L
Ro"'JJ~'[e~p' - e~p'(T)J
(5.392)
~JJ'
where, for l!fJIP(T), we use either (5.388) or (5.389), depending on the circumstances. In order to make (5.392) more concrete, let us suppose we have a two-level system. Thus we may have a spin particle quantized by a static field in the z-direcrion. We label the states 1 and 2 and find that the density matrix has four elements Ull> 022, un, U'2t· In the relaxation tenns Ra~ ,/Jll', the only tenns of imponance (as we have seen) involve cr - crl = fJ - If. The only tenns Ihal count are therefore
!
R Il ,22 =
R22,1I
R12,12 = R21,21
== I/TI == - 1/1"2
(5.393)
By assuming 1i2 joins states 1 and 2 only, and denoting (l11i2(t)12) by 1i12(t), we find dUl1
_'_e-1{,/"·
,
dM dt
--=/
d.
dU22
=-dt" =
t!22 - ",1 -
[m(T) -
ell (T)] + ii; ( UI2 ~ (t) 1L21
TI
del2
Ul2
-- = -~
&
i
l
i
+ -(£2 - £1)UI2 + -:-(U11 - U22)1i 12(O h h
~ (t)n )
- Iq2
c:21
(5.394)
(5.395)
If E2 is larger than £1 and '}-{12 oscillates at frequency w, we can solve (5.394) and (5.395) in steady state by assuming that
="21e-i~t
l!12 = r'12ei~t,
.!?21
el1 = rll ,
l!22 = r22
(5.396)
where the roo, 's are complex constants. The details of the solution are left as a problem. but the form of the answer is identical to that of the Bloch equations.
'17
If we write
'H2(t) = V cos wt
(5.397)
6. Spin Temperatnre in Magnetism and in Magnetic Resonance
where V is an operator, and define Wo by the relation E 2 - El == hwo, we have in the limit of small V (that is, no saturation) e-El/ kT
1 -El/kT 1'" = ell(T) = e EI/kT + e E21kT ';::' -e 2
'"
_ _ V"T2 [(T) - 2'n 1·( ) ell +IW WOT2 S! IT2 V I 2W O I + i(w WO)T2 4kT
e22
(1')J
6.1 Introduction (5.398)
We note that 1'12 differs from zero only near to resonance and that T2 characterizes the width of frequency over which 1'12 is nonzero. If the states 1 and 2 are the two Zeeman states of a spin ~ nucleus and a static magnetic field parallel to the z-direction, then the transverse magnetization Alz has matrix elements only between states I and 2, the diagonal elements being zero. Therefore (M:r(t»
=
1'12eiwt(2IMzll) + 1'2Ie-iwt(lIMzI2)
= 2 Re {TI2eiwt(2IMzlI)}
(5.399)
Taking (5.400)
V = -MzHzo and recalling that X is defined as
(5.401)
(Mz(t)} = Re {xHzoeiwt} we see that
wo!(IIM.12)I'
'T2
X(w) = 1 + i(w _
2kT
WO)T2
(5.402)
Now, using the fact that I =~. we have X(w) =
iT2
1 + i(w -
WO)T2
wO'Y 2 1i. 2 ](l+I) -
2
3kT
(5.403)
which agrees with the expression for the Bloch equation derived in Chapter 2. We note that we could detennine both 71 and T2 from first principles by computing R",22 and R I 2,12' Alternatively we could simply treat Tl and 72 as phenomenological constants to be given by experiment. If there are more than two levels to a system, the solution may be carried out analogously by simply setting all off-diagonal elements of gaol equal to zero, except those near resonance with the alternating frequency (Err - Eo, S! liw).
218
In Chapter 5 we employed the concept of spin temperature to discuss relaxation. The idea of spin temperature was introduced by Casimir and du Pre [6.1] to give a thermodynamic treatment of the experiments of Gorter and his students on paramagnetic relaxation. It was Vall Vleck [6.2] who first employed the concept for a detailed statistical mechanical calculation of the relaxation times of paramagnetic ions. 80th in this case, and also in his general statistical mechanical treatment of static propenies of paramagnetic atoms [6.3], he recognized and emphasized the fact that expansion of the partition function Z in powers of lIT enabled one to calculate Z without the necessity of solving for the energies and eigenfunctions of the Hamiltonian. Waller evidently was the first person 10 use this property [6.4]. From the partition function, one can compute all the static properties of the system, such as the specific heat. the entropy. the magnetization, and the energy. For example, the average energy of a system. E, at a temperature T is given by
E = kT' :T In Z
(6.1)
In 1955. Redfield [6.5] showed that the conventional theory of saturation did not properly account for the experimental facts of nuclear resonance in solids. In one of the most important papers ever written on magnetic resonance. he showed that the conventional approach in essence violated the second law of thennodynamics. He went on 10 show that saturation in solids can be described simply by applying the concept of spin temperature 10 the reference frame that rotates in step with the alternating field H \. To understand his ideas, one needs to understand certain concepts which predate the discovery of magnetic resonance - ideas such as adiabatic demagnetization. In this chapter we begin by describing a simple experiment which displays the failing of the pre-Redfield theory of magnetic resonance. Then we tum to a discussion of the use of spin temperature in nonresonance cases to build background for the application of these same ideas in the rotating reference frame. We then discuss the Redfield theory of saturution in solids.
219
6.2 A Prediction from the Bloch Equations Let us consider a simple resonance experiment with a rotating magnelic field of angular frequency w. lI"ansverse to the stalic field B o• tuned exactly 10 resonance w: ,Ho
(6.2)
We discuss it by means of the Bloch equations. It is convenient 10 transfonn to a reference frame rotating at w with HI along the x-axis as is done in Sect 2.8. Exactly at resonance. the Bloch equations become dM~
----;It "" -..,MyH I + dM~
Mo - M~
(6.3a)
Tl
Jkrz
(6.3b)
-;U""-T2
dMy "" -vM HI _
dt
,:
~
(6.3c)
T2
Suppose we now orient M along HI sO that at t "" 0 Mz "" Mo. Jvly "" M: "" O. From (6.3b) we see that M z will decay 10 zero in a time T2 • The low HI steadystate solution of the Bloch equations shows Ihat they describe a Lorenlzian line with a frequency widlh I &:-
.~
T,
,
For solids typically Llt..,)
';!'
'1Hnei,hbor ~ '11l ~ '1.-../1(/ + I)/Z J (tJ
a
6.3 The Concept of Spin Temperature in the Laboratory Frame in the Absence of Alternating Magnetic Fields Let us now tum to a discussion of the application of the concept of spin temperature to magnetic experiments not involving resonance. A typical system we might consider is a group of N spins of spin I, gyromagnetic ratio '1, acted on by an external field Ho, and coupled together by a magnetic dipolar interaction represented by a dipolar Hamihonian Hd' We denote the Zeeman Hamiltonian by Hz. The solutions of the Schmdinger equation are then wave functions tPn of energy Ell of the total system.
'li",,: ('liz + 'lid)"" : E"""
(6.5)
(6.6)
Unfonunately, (6.6) is exceedingly difficult to solve, depending as it does on the coordinates of 10'22 spins. One can assume, however, that if the spin system is in thennal equilibrium with a reservoir of temperature 8, the various states n of the total system would be occupied with fractional probabilities PII given by the Boltzmann factor I'll ""
where H ~gl . I bo r is the nuclear magnetic dipole .field due to neighbors, and a is . . the distance to the Z nearest neighbor. For typical solids Llt..,) IS of order of a few tens of kilocycles (e.g. .t1w S! 211'" X lO kc for AI metal). Funher examination of (6.3a) and (6.3c) shows that they do not involve M~ and that if M and M z are initially zero they would remain zero were it not for the tenn invofving T t . If T} ::> T2, therefore, and for times up to about T I after M has been orienled along H I, we still have My "" M: "" O. Therefore these eqlw.tions predict that wilen M is aligned along HI, it wUl decay to zero in a time T2, typically of order 10- 4 to lO-5 sec. Experimentally this prediction (decay in T2 when M is along H) is found to be correct for liquids, but it is not correct for solids. Rather, for solids it is found that as long as HI is turned on and is sufficiently strong, the decay rate of M~ is much more like T, than a time T2 which characterizes the line width. Redfield first stated this fact on the basis of his steady-state experiments, but he did not actually perfonn lhe experiment we have described. That experiment was first performed by Holton et a!' [6.6]. Why do the Bloch equations fail for 220
solids but not for liquids? Redfield has given an explanation. We will discuss the conditions for validity of the Bloch equations later in this chapter, after we have discussed some important background material on spin temperature in cases where there is no alternating field HI present.
E,,/k8 .!..eZ
(6.7)
where Z is the panition function (6.8)
" Quantities such as the average energy E and magnetization M ~ would then be given by
E:E~&
~~
"
(6.9b)
"
As we have remarked, Van Vleck recognized thai expressions such as (6.9) could be evaluated wilhout solving the Schmdinger equation because they could be expressed as traces. For example we can express the panition function as a lrace as follows. (6.10)
"
"
Since the trace is independent of lhe panicular representation used
10
evaluate it. 221
we can use a convenient representation. For example. we could evaluate (6.10)
in principle by using as basis functions the eigenfunctions of the z-component of spin Izk of all the individual nuclei. However, to do so, we need to lake one more step, expansion of Z in a power series. It is often valid for nuclei and for eleclTons to use the high temperature approximation. We expand the exponential in a power series, keeping only the leading lenns. Then the sums are easy to do. 2 Hl1i N', } Z=Tr { 1-/.;8+2/.;28 2 -'" =(21+1) +2k26 2Tr {'Jt}+··(6.11) where we have used the faci that Tr {1t} = 0, as can be readily verified for both
Hz and 'HdUsing these methods one finds C(H2 + lJ2) 0 L
E=
where
8
N"f2fI 2I(I + 1)
C=
(6.12)
(6.13)
3k
is the Curie conslant, and HL is a quantity we call the local field, which is of the order of the field one nucleus produces at a neighbor (several Gauss) and is defined by (6.14) 1 T< {H'} L- k(2I+l)N d Since the trace in (6.14) can be computed, HL may be considered to be precisely known. One finds H[:: -y 2r!2[(I + I) ~)1/rjk)6 (6.15)
CH' -
j
One can compute the magnetization M and finds it obeys Curie's law
M= CHo 8
q=
E+kBlnZ
(6.17)
8
Evaluating Z and E we get C (H 2 +H 2) ~::Nkln(2I+I)-0 L
2
8'
(6.18)
6.4 Adiabatic and Sudden Changes The significance of these results is more fully realized by considering the behavior of the spin system when the applied field Ho is made a function of time. For simplicity, let us assume the spin system is thennally isolated from the outside world and that it mayor may not be in [hennal equilibrium. The first of these assumptions is satisfied if the experiments we conduct are perfonned on a rime scale short compared to the spin-lauice relaxation time. The assumption implies that the Hamiltonian of the system includes only variables intemal to the system since spin-lanice relaxation results from tenns involving variables both intemal and external 10 the system as illustrated by (5.33). If the system is in internal themlal equilibrium, a spin temperature applies, so that (6.18) holds. We can then consider three cases.
(6.16)
Note that this is a vector equation, so that M and Ho are parallel. A moment's reflection shows that (6.16) is truly remarkable. It states that the vectors M and Ho are parallel no matter what the size of Ho as long as the high temperawre approximation is valid. Suppose Ho is small, comparable to the local field a nucleus experiences from its neighbors. One might then suppose that nuclei would tend to line up along the direction of the local field, not along Ho. It would seem reasonable to suppose that the degree of polarization one could achieve per unit of applied field would be less when Ho «..H L than when Ho:» H L . Equation (6.16) shows Ihat this intuitive argument is incorrect - the degree of polarization per unit of applied field is independent of the size of Ho relative to the local field. Not only is this true of the magnitude of M, but also of the direction as well. 222
Another useful property of (6.16) to note is that when Ho :: 0, M :: O. Thus, suppose Ho were tumed 10 zero so suddenly that M does not have time to change. Immediately after Ho :: 0 we have a case where M f. 0, Ho :: O. But if there..were a temperature, (6.16) shows M must be zero. Therefore we can use (6.16) to conclude that at this instant of time the system is not describable by a temperature. Another quantity of great utility is the entropy ~. We know from statistical mechanics that Ihe entropy measures the degree of order in a system. In a reversible process in which there is no heat flow into or out of a system, the entropy of that system remains constant.
a) The Hamiltonian is independent of time (the applied field is static). In this case, the average energy is constant in time, whether or not the system is describable by a spin temperalure. If the system has many parts which are strongly coupled together, but is not initially in a state of internal equilibrium, we expect that irreversible processes within the system will eventually bring the system into an internal equilibrium describable by a spin temperature B. During that process, the energy is conserved since a time-independent Hamiltonian corresponds 10 a system on which Ihere are no applied forces. Proof of the constancy of the energy is left as a homework problem (Problem 6.5). b) The Hamiltonjan changes slowly jn time. The criterion of slowness is that at all times internal transfers of energy shall be fast enough so that the system is always describable by a single temperature B. Under this circumstance, Ihe changes are reversible, and the entropy of the system remains constant. 223
system. The condition will be satisfied if we make the changes on a rime scale rapid compared to the spin-lattice relaxation time T!. Frequently one has T1's of seconds, and by cooling may achieve Tl'S of hours. Such long limes are practically infinite. The second condition we must satisfy is that after each small change in Ho we must allow a new temperature to be reached before making another small change. This condition is typically that we must change Ho slowly on a time scale defined by the precession period of nuclei in the local field of neighbors (I/-yHL)' This time scale is a few tenths of a millisecond. Between these time intervals there is a readily achievable range for which Ho can be changed adiabatically. For an adiabatic change we have a constant entropy. Thus from (6.18) we get that (HJ+HVI02 remains constant. If we stan in an initial field Hi at temperature 0i and change the field adiabatically to a final value Hr, the temperature Or is given by the relation H.I2 +HI.2 _ H2r +H2). (6.23)
c) The Hamiltonian changes discontinuously in time. Such a change could occur if Ho can be changed quickly. The term "discontinuous" means that the change is so fast that the various spins making up the system do not change direction during the change. Let us now investigate these various cases more fully. I) Time-Independent Hamiltonian
Consider a system not in thermal equilibrium initially. Let us suppose that the parts of the system are coupled together. We then expect that eventually the system will achieve an internal equilibrium described by a "final" temperature If we know the energy of the system at t = 0 (call it Eo), we can compute I~e temperature Or as follows, making use of the fact that for a time-independent Hamiltonian the energy is conserved. _ . Utilizing (6.12) which relates the energy E to the temperature, and applymg conservation of energy, we get that
o.
E=Eo
or
8r -
(6.\9)
C(H6 +Hl) Sf = Eo
A famous result, cooling by adiabatic demagnetization, can be seen in (6.23) by taking and <: corresponding 10 what happens when a sample initially in a strong magnetic field has that field turned to zero. Then we
H;:> Hl,
(6.20)
eRo
8r H L -=-<:\ 8i Hi
(6.2\)
where we take Of from (6.20). During the process of establishing internal thermal equilibrium, the entropy is not constant since irreversible processes are taking place. But eventually the entropy O'r is given by the thermal equilibrium value for a temperature Or C(H 2 +H2) ~02 I.
(6.24)
Thus, the final temperature of the spins is much colder than the initial tempcrnture. If initially the spins were in thennal equilibrium with a thermal bath (such as the lauice cooled to liquid helium temperatures), the final temperature would be a good deal less than the bath temperature. Note that the bigger Hi and the smaller H L, the greater the cooling. To reduce HL' it is common to dilute the magnetic atoms. Figure 6.1 shows Or versus Hr for an initial field Hi much larger than the local field, and for an arbitrary initial temperature 0i' Note that as one lowers the
Sf
O'r= Nkln(2I+ l)-
Hi Ht
find
We can compute the magnetization Mr which finally results after internal thermal equilibrium is reached by means of Curie's law
Mr=--
Of
(6.22)
f
If the Hamiltonian can be divided into two pans which commute, the energy of each pan is separately conserved. The system then cannot be expected to reach an equilibrium described by a single temperature, but rather is expected to reach an equilibrium in which each commuting pan is described by a temperature. We encounter this situation when we apply these ideas to the rotating frame later in the chapter. In some cases the subsystem may not have enough complexity to make us confident it will eventually be describable by a temperature, but even so we ohen blithely proceed to assume that a temperature is achieved.
0,
2) Slow or Adiabatic Clwnges
We keep in mind that to be adiabatic, a change in Ho must satisfy two conditions. The first is that there should be no heat flow into or out of the spin 224
",
Fig.6.1. The final temperalure 6 f reached al a final magnetic field I1f ror adiabatic changes in the applied field rrom an initial value ilL at tcml>erature 8;. Noll' lhat 8r does not chllnge much ...·ith IIr for IIr < II.. 225
field from its initial value. the temperature drops until the applied field becomes comparable to the local field. Funher decreases in applied field do not then produce much lowering of the temperature. The magnetization can be computed from Curie's law if the temperature is obtained from (6.23). For the case that there is an initial magnetizmion Mj in an initial field Hi much larger than HL. the final magnetization M r is then easily shown 10 be Hr Mr = Mi (6.25) 2 V/H r +H2L
This result is shown in Fig. 6.2. NOIe that during me cooling process, M r remains at the initial value Mj until Hr gets comparable [0 the local field HL. because for large values of H. 9r is propor1ional to Hr.
thennal equilibrium between the spins and lattice in the field Bo. If Yo:» H L• the magnetization is small, but nevenheless it is not zero. One of the remarkable features of (6.25) or Fig. 6.2 is that when H f = 0, Mr = O. As we have remarked, this is a very general consequence of Curie's law no matter what the temperature. (Note, however, that the derivation of Curie's law utilizes the high temperature approximation. At low enough temperatures this approximation breaks down, allowing spontaneous magnetization to be a possible zero field state). What is remarkable is that the degree of order of the spin system is just the same when M = 0 as when M = Mi. Although it is clear that a system with net magnetization is ordered, how can there be order when AI = 01 The answer to this paradox is that even when H 0 = 0, spins still experience magnetic fields owing to the presence of their neighbors. A typical spin will point either with or against the local field it experiences. For a highly ordered system, there will be a substantial excess pointing with the local field rather than against it Since the local fields at different nuclear sites have ran· dom orientation (we here rule out highly ordered spin arrangements such as in a ferromagnet), there is no resultant muaoscopic magnetization resulting from the alignment along the microscopic local fields.
3) Sudden Switching
Fig. 6.2. MlIgno:ti:tation AI, versus "pplicd fidd IIf for adiabatic dl"ngcs in IIf. It is &lI5lImed that III > II~. Noto: thllt M, is independent of 'II{ for II~ > and goes to zero when IIr goes to :tetO
We have considered a process which is reversible. Now we tum to one where things happCn suddenly, resulting in irreversibility. Suppose we describe the system by a wave function t/J. TIle time-dependent Schrtxl.inger equation is
h
-i
a~ {It = 7i(l)~
11£,
II,
This equation is ploued in Fig.6.2. Note that if we lower the field until it is zero, M f = O. However, since the field changes are reversible (and in fact at all times keep the entropy constant), we can recover Mi by raising Hr from zero back up to its initial value Hi. The recovery of Mr in such a process d~s 1I0t involve spin-lattice relaxation since, as we have postulated, everything is done on a time scale shon compared to Tt! This curve is often a surprise to those of us who first learned magnetism by studying magnetic resonance, since we learned that one needs TJ to produce magnetization from an unmagnetized sample. Actually, if one uses Curie's law and (6.23), one sees that if one stans in zero field with spins which are in thennal equilibrium with the lauice (8 = ()/), the mere act of adiabatically turning on the static field to H o ,. HI. will produce a magnetization. The size is (6.26) where Mo is the magnetization one gets when the spin-lattice relaxation produces 226
= -M·H(l)~+7id'"
(6.27)
where M is the tOlal magnetic moment operator and 1id the dipolar coupling. The time dependence of 1i arises because the applied field H is time dependent. For the case of sudden switching, we take H as independent of time except for t = 0, at which time it jumps discontinuously from one value to another. Denoting by 0- and 0+ times just before and just after t = 0, we then have
~(O+) -
0+
"'(0-) = -
J*7i(l)~(l)dl
=0
or
(6.28)
0-
(6.29) since 1i(t), though discontinuous, is nevenheless never infinite. Thus we have that the wave function just before the switch is identical to its value juS! after. We can utilize (6.29) to see that a sudden change in H produces a change in the expectation value of the energy E. The expectation value of energy at any time t, E(t), is E(l) = (~(l). 7i(l)~(l))
(6.30) 227
Thus
whence we gel
E(O+) = -C H(O-) . H(O+) _ C H~
E(O-) = (,,(0-). 'II(O-),,(O-» = -(M(O-» . H(O-) + ('IId(O-»
8.
(6.32)
(6.33)
But
is conserved. 1Oerefore.
Ukewise
(6.41)
The fact that "'(0-) ::: "'(0+), however, means that the expectation values of both magnelization and dipolar energy are the same at t ::: 0+ as at t ::: 0-. This resuh expresses Ihe physical fact Ihal all spins point the same direction al t ::: 0+ as they did at t ::: 0-. Thus we can write
(6.34) This equation is very useful since in general the system just before Ihe sudden change is assumed 10 be in internal thennal equilibrium with temperature OJ. Thus we can compute the expectation values by the methods of Sect. 6.3. So far we have nor employed the density matrix notation, simply to make Ihis chapler accessible to those readers who have not yet become familiar with it. We can write our results compactly by recognizing that (6.29) is equivalent to saying that lhe density matrix f! obeys the relation
(6.35) Thcn
(6.36) Assuming thermal equilibrium at t ::: 0- at a lemperalUre Bi we get
_
riCO ) =
cxp(-'II(O-YkO,) Z(O )
(6.37)
Therefore
('IId(O-» = Tr {e(O-)'/{d} Tr {'lid cxp (-'II(O-)lkO,») Z(O-)
(6.38)
Manipulation, using the high temperature approximalion plus the definition (6.14), gives
('IId(O-» = 228
(6.40)
As discussed earlier, immediately after switching the field. the spin system is in general not in internal thermal equilibrium even though it was in thennal equilibrium at t = 0-. If we wait a long enough time, we expect a temperature to be achieved. Call that time if and the temperature 8r- We can compute 8 r by recognizing that fOT t > 0 the Hamiltonian is lime independent, hence the energy
where (M(O-» and (1{d(O-» are the expectation values of magnetization and dipolar energy at t = O~. defined as
(M(O-» " (,,(0-). M,,(O-» ('IId(O-» " (,,(0-). 'IId"(O-»
8j
(6.31)
CH'L 0,
(6.39)
E(tr) = _CIH'(O+) + HlJ
0, and E(O+) is given by (6.39). Thus
H(O+)' + H' 8r - (J. " - 'H(O ).H(O+)+Hl .
(6 42) .
The significance of (6.42) can be seen by considering a particular example in which the applied magnetic field is turned suddenly from its initial value. Ho, to zero at t = O. Then H(O-) ::: Ho. H(O+) ::: O. giving
Or = 0,
.
(6.43)
We contrast this with the result of turning the field slowly 10 zero
HL
Or::: OJ Ho
(6.44)
The sudden tum-off leaves lhe temperatu~ unchanged, the slow tum-off produces cooling. The results of sudden changes are summarized in Fig. 6.3. for Ihe case described above in which Ho is lurned suddenly to zero at t ::: O. Adiabatic and sudden changes in Ho have been very useful in magnetic resonance. One of the first uses of adiabalic changes was to measure the magnetic field dependence of the spin-lattice relaxation time. In 1948, Turner el al. [6.7] were studying [he spin-lattice relaxation time TI of protons in insulators. The relaxation times were, in some inslances, many minutes long. To observe Ihe dependence of T 1 on stalic field, they used a field cycling lechnique in which they observed the resonance at a high field. bUI allowed the spins to relax. in lower fields to which they cycled belween their observations. Pourui [6.8] discovered that the nuclear relaxation times in a crystal of LiF were so long thai he could take his sample out of the apparatus into Ihe earth's magnetic field and then return it with only a small loss in magnetization. With Ramsty [6.9] he demonstrated that if this sample were removed from the strong field 10 a small static field (40 Gauss
'29
III
i
/ P 'V!v ~
£1
Fig. 6.3. Magnetic field (11), wave function (l{!), energy (E), spin temperature (0), and rnagnc~ization (M) as functions of lime for the case of a magnelic field turned suddenly to zero
i - - - - - !-
,
, ,,
~N;;O~ _ _ {%
,
I
,
"I
-----'--"-----
or less), subjected to an audio-frequency magnetic field, and then returned to the original magnetic field, he could detect aI the strong field the resonant absorption of energy by the spin system when in the weak field. The experiment we have just described is closely related to the famous experiment on negative temperarures by Pound and Purcell [6.10]. They 100 started with the LiF sample in a strong field, but removed {he sample to a solenoid whose field they suddenly reversed, producing a situation in which the magnetization is anliparallel to the static field. They then returned the sample to the strong field where they observed the relaxation of the magnetiz.1tion component along the static field. They point out that such a circumstance, with the upper energy Zeeman levels more populated than the lower, corresponds to a negative Zeeman temperature. Any system with an upper bound on its energy levels can in principle have a negative temperature. In addition to the original paper by Pound and Purcell, we call the reader's attention to the wonderful account of this experiment and discussion of the negative temperature concept to be found in Van Vleck's [6.11] lecture on the concept of temperature in magnetism. (Professor Van Vleck cautions that there are some incorrect numbers in his paper.) This imponant experiment in which a population inversion was produced is the forerunner of the maser and the laser. Demagnetizing to zero field is the basis for the experimenls of Hebel and Slichter [6.12] and Redfield and Anderson [6.13,14] to measure the nuclear relaxation in superconductors. The problem here is that since superconduclOrs 230
exclude the magnetic field, it is hard to observe a magnetic resonance in a metal in the superconducting state. One starts the field cycle with a magnetic field of sufficient strength to suppress the superconductivity. When one turns the field to zero, one achieves two effects: (1) the nuclear spins are cooled and (2) the sample becomes superconducting. With the field zero, the nuclear spin temperature relaxed towards the lattice temperature so that when the magnet was turned on again after a variable time T, the metal returned to the nOTlllal state and the spins warmed to a holler temperature than they had before the cycle began. The increase in temperature is measured by a swift pass though resonance. By varying T, one could deduce the zero-field spin-lattice relaxation time. The NMR results showed that as one lowered the sample temperature below the superconducting transition temperature, Te , the nuclear relaxation rate was at first faster than it would have been if the metal were normal, then at lower temperature it became slower. The temperature dependence of the nuclear relaxation rate is dramatically different from the temperature dependence of ultrasonic absorption, as was shown by Morse and Boltm [6.15]. They found that ultrasonic absorption dropped rapidly relative to its value in the normal metal on cooling below Te · The contrast between the behavior of two low energy scattering processes, nuclear relaxation and sound absorption, is difficult to understand in a one-electron theory of metals, but finds a natural explanation in the Bardeen, Cooper, and Schrieffer (BCS) theory of superconductivity. These experiments constitute a direct verification of the concept of eleclTon pairs, which is the basis of the BCS theory. (See, for example, Leon Cooper's Nobel Prize lecture [6.16].) Anderson [6.17] working with Redfield, combined field cycling with the application of an audio-frequency magnetic field applied while the static field was zero to heat the spins to plot Ollt the zero-field absorption characteristics of a spin system. Thus they used field cycling to give them the sensitivity of resonance in a strong field to monitor the effects they produced in zero field. Abragam and Proctor [6.18] did further studies establishing the validity of the spin temperature concept, again using LiF. An important result of their experiments was their observation of the transfer of energy between the twO spin systems (Li 7 and F19) even in a static field which greatly exceeded the dipolar fields exerted by the nuclei on their neighbors. We discuss these experiments further in Sect. 7.10.
6,5 Magnetic Resonance and Saturation The analysis of magnetic resonance by Bloembergen et lIl. using standard perturbation theory is given rather compactly in (1.32) - the differential equation for the population difference, n, between the two energy levels of a system of spin particles,
4
dn
no-n
- '" -2W(w)n + - dt T1
(6.45) 231
W(w) is the probability/second that a spin will be flipped by the radio-frequency field H t . Standard perturbation theory shows W(w) = i-y2 Hrg(w)
(6.46)
where g(w) is a funclion nonnalized to unit area having the same dependence on frequency as the absorption line -that is, it expresses the fact that the frequency of HI must be close to resonance for H, to induce transitions. no is the thennal equilibrium population difference, and TI the spin-lattice relaxation time. It is always JX)ssible, at least conceptually, to consider TI infinite, in which case (6.45) is especially simple to solve. n = Ae- 2Wf (6.47) where A is a constant of imegration. It is well to recall the conditions for the validity of (6.46). 1bey are two: i) ii)
The perturbation matrix elements inducing transitions must be small compared to the width of the final state energy levels. This means HI < H L. The wave function must not change much. We note, however, that (6.47) predicts that fl --f 0 as t _ 00.
To satisfy condition (ii) we expect that we must consider times less than 1/1V. Thus, though it is always easy to meet condition (i) by making H, small, no matter how weak HI, if we wait long enough we violate (ii). [Note we are requiring here also that I/W < TI, otherwise the Tl tenn would rescue condition (ii) even for times long compared to I/Wj. We have the interesting problem, therefore, that we do not know how to integrate the equations of motion beyond a time for which n is almost its value at t = O. (See Fig. 6.4). The solution of this problem was found by Redfield [6.19] in a truly remarkable paper, the more so when one recognizes that it was his first work on magnetic resonance. In it he shows that the Bloch equations, when applied to a solid, violate the second law of thennodynamics. The essence of his approach is to note that a resonant time-dependent perturbation, no matter how weak,
".
of l"lliitiity of ~Rrgiolt com'elJliOIl(lI/H:1(llrlJo/io/f /lleor)'
will eventually produce large effects. Whenever a perturbation of small size can produce a big effect, it is dangerous to treat it lightly. He therefore eliminates the time dependence by transfonning to a reference system in which the Hamiltonian..is essentially time independent. The residual time dependence is not of the dangerous variety. For such a transfonned system, the energy is conserved. Moreover, the system is highly complex, consisting of many interacting spins. One can thus predict that after a sufficiently long time the system will be found in a state of internal equilibrium. That is, it will be in one of its most probable states. Phrased alternatively, the energy states will be occupied according to a Boltzmann distribution at some temperature 8. We consider the Hamiltonian 1t, given by
1/ = 1/z(I) + 1/d
where llz(t) is the Zeeman interaction with the static field Ho and the rotating field of amplitude HI and angular frequency w rotating in the sense of the nuclear precession. "The TOtating field makes 'Hz time dependent. We are, of course, considering T I infinite. Utiliz.ing the methods of Sect.2.6, we transfonn to the rotating reference frame, gelling a transfonned Hamiltonian J(:
'It = -~h(Ho-wh)I.+HII.J+1f.: + term oscillating at ± w, ± 2w -Iz sin? + I y cos; = e- i/ .? Iyei !.?
and
(6.50) (6.51)
to transform products such as ei!,w.l IzIye -iI
.w. t
(6.52)
to expressions involving sin Wit and cos Wit. The term 'H~ is that part of the dipolar coupling which commutes with Ii' Physically, it is the part of 1id which is unchanged by rotation about the z-axis. (The two statements are of course equivalent since one can consider Ii as generating rotations). II is the sum of the tenns A and B discussed in Sect. 3.2. The fonn of 1t~ is
o
"·ig.6.4. Conventional 6aturation theory predicts that the I>opuiation JI goes to zero exponentially with lime t. However, the assumptions on which il is based are valid only for the initial PlITt of the curve where n s:' no
(6.49)
In deriving (6.49) we have used relations such as
"(2/i,2
1id = - 4
"
(6.48)
L
j,k
(1-3cos 2 8j1.) :I r jk
(3I:jI z k - Ij'
h)
(6.53)
where 9jk and rjk are coordinates of nucleus j with respect to nucleus k. The term in the square brackets in (6.49) may be considered as the coupling of Ihe spins to an effective static field He as we noted in Chap. 2.
He
=
k(Ho -w/-y)+iH I
(6.54)
2/W 232
233
In the absence of the time-dependent tenns, the energy levels of 'HI would be split by He·and the dipolar couplings, so that we expect typical splittings to be IJE ~ iliJHi + (6.55)
M=C
HJ
(6.56)
Now, in the absence of HJ, 'Hz and 'H3 commute, since the fact that
H3
[Iz , 11:3] = 0 was the definition of1f.3. Under this circumstance 'Hz and would separately be constants of the motion. However, if HI f. 0, ['Hz, H~] f. 0, and the Zeeman and dipolar systems can then exchange energy. Since H is independent of time, the total energy is conserved. Moreover, the system is very complex. Redfield therefore postulates that no matter what the state of the system at t = 0, a long time later it will be in a state of internal equilibriulll described by a Boltzmann distribution. In other words, there will eventually be a temperature 6 which can be assigned to the spins. We can thus say that the density matrix g is
e=
..
e-'H/k9
(6.57) Z with Z = Tr {exp(-H/k8)}, where 'H is the effective Hamiltonian of (6.56). Of course, we expect that after a long enough time Redfield's hypothesis would be fulfilled (unless there were some hidden selection rule which we have overlooked, are perfectly isolated from one llnother if HI such as the fact that 1f.z and is zero). But the really important question becomes, how long does it take to reach equilibrium? The answer to this question clearly depends on the size of HI, since HI is needed to prevent isolation of the dipolar and Zeeman systems. We return to the question later, but for the present consider that the time is short enough to make the establishment of a temperature practical.
H?t
6.6 Redfield Theory Neglecting Lattice Coupling
2
CH,2 _
I T, «'lj0)2) L- k (2I+I)N d
E = _C(H; + H'I> 8 234
(6.58)
(6.60)
(6.61)
Evaluation of the trace of (6.61) gives, for a system with only one species, (6.62)
where {IJH 2 } is the second moment of the resonance line. Following our earlier convemion of omitting primes. we shall now use H L for H~, using the prime only when we wish to distinguish between the local field in the laboratory reference frame and the rotating reference frame. It is important to notice that the Redfield assumption leads to Curie's law and that the vector nature of the law shows that the nuclear magnetization is always parallel to the effective field when the system is describable by a temperature in the rotating frame. Thus, if one is exactly at resonance, where He = iHI' the magnetization is perpendicular to the static field. Since the fonn of (6.58), (6.59) and (6.60) is identical to that of the corresponding equations in the laboratory frame, most of the equations of Sect. 6.4 can be immediately applied to the rotating frame by repl:lcing Ho with He and with
HE
Ht[.
6.6.1 Adiabatic Demagnetization in the Rotating Frame An adiabatic demagnetization in the rotating frame can be perfonned readily, as was demonstrated by Holron [6.20]. Suppose that initially HI = 0, and the sample is magnetized 10 Its thermal equilibrium value kMo. Let us shift Ho far above the resonance value wI'"'(, and tum on HI' C'Ne assume Ho - wI'"'( to be much bigger than HL and HI.) We now have an effective field which is virtually parallel to M. Next we change Ho to approach resonance at 11 rate sufficiently slow to satisfy the criterion for a reversible change. C'Ne are here assuming T l to be infinite, which we achieve in practice by perfonning all experiments in a time shorter than T j ). We then have that
M = Mo The significance of (6.56) can be appreciated by calculating again the energy E, entropy a, and magnetization M. We find easily that
(P
where C is the Curie constant, and where
HE H;,
1f. = -'"'(hI· He + 1f.3 = 1f. z + 1f.~
(6.59)
a=Nkln(2I+I)- C (H;+HII)
HE
where the square root is a convenient way of including the two limiting cases of He>HL and He«H L . The time-dependent tenns will connect states of order iliHo apan in energy. Unless ~ + a very low resonance field indeed, the time-dependent tenns are far from resonance and can be neglected since they are unable to produce transitions. They are not dangerous. We thus obtain a Hamiltonian which we call 1f., omining primes for simplicity of notation.
He 8
He JHi+ H [
(6.63)
Notice that M is parallel to He' as is required by Curie's law. Thus, as one approaches resonance, M changes direction, always pointing along He. In general, the magnitude of M also changes, unless Hi ~ HE. We can experimentally measure M by suddenly turning off HI, leaving M}o precess freely about 235
Ho. The induced voltage immediately after turn-off is proportional 10 M',t.. (Use of a phase sensitive detector enables one to measure Al'1: and 1L1y , but for these experiments My = 0). One can measure M z by noting Ihm though M'I: decays to zero within a time of order I/-yH L, after tuming off HI, M l does not change. One can thus wait till M',t. has decayed, and then apply a 7r/2 pulse which rotates M z into the x-y plane for inspection. The theoretical values of M'I: and M l are HI
HI
MQ
------------ ----- -----------------_.-
(6.64)
M. = M - = MOr~==;; He ./H2 +H2
vel,
ho ho M, = M - = MO'f~~=,;' He I H 2+H2 V el.
(6.65)
where ho is the component of the effective field in the z-direction:
110
== (Ho - w/-y)
(6.66)
Notice that if one does an adiabatic demagnetization exactly to resonance, the value of magnetization ~=~
HI /H2 +H2
V
I
0~
L
will persist indefinitely as long as HI remains on (actually, when relaxation to the lattice is included, it decays, but on a time scale typically of order T I ). We contrast this prediction with our earlier conclusion from the Bloch equations that M:z; decays to zero in T2 , where T2 ~ I/-yH1•. The factlhat M does not shrink as long as He is kept constant, and that M precesses in step with HI is often is comparable 10 M:z; described by the graphic tenn "spin locking". If will be less than Mo. However, the "loss" of magnetization is recoverable. Were one 10 go back off resonance, M would grow back to M o when H;:» Figure 6.5 shows (6.67). When Redfield proposed his theory. the fact Ihat the spins were locked to He was one of the most surprising results. It is, of course, nOlhing but the rotating frame equivalent of the statement that the magnetization in the usual laboratory frame adiabatic demagnelization has a one-to-one correspondence with Ho, as expressed in (6.25). Note that if one pulses on Ho when (Ho -w/-y):»HI and H t • and then changes H o so thai one passes through resonance, continuing until one is far on the other side of the resonance (Ho - w/-y negative), one will have turned M antiparallel to Ho. Moreover, although near to resonance, one might have M < Mo (if HI < Hd; by the time one is far from resonance one would have M '";;£ Mo· This experiment provides a simple means of lllrning over the mngnetization. One can see from (6.64) that if one demagnetizes exactly to resonance, the magnetization will be the full Mo provided 1I I :» HI.. TIle resonance signal is then as big as can be achieved using a 1r/2 pulse. If HI ~ HI,. one will not
lIf
El,
Hl.
236
fl i . fI, )'ig.6.S. The transverse nmgnetization At" produccJ in a demllgnetization experiment in the rotating frame, as a runction of the strength or alternating field If, employed. Ih is assumed to be turned on when If a is we11 off resonance, with the slImple initially at its thermal equilibrium magnetization {\'fa
achieve the full magnetization at resonance. Were one to reduce HI slowly to zero after arriving at resonance, M would shrink to zero. In this manner all of the order represented by Mo prior to demagnetization would have been put into the dipolar system, thnt is, into alignment of spins along their local fields. One could, at a later time, slowly turn H t back on again, thereby recovering the magnetization. 6.6.2 Sudden Pulsing A situation frequently encountered in experiment and interesting to contrast with adiabatic demagnetization is the effect of a sudden change in He. We treat an especially simple case, that of suddenly turning on HI at time t = O. We assume that before we tum on HI, we are off resonance by an amount ho, with the system initially magnetized to Mo along Ho. We utilize the ideas of sudden changes discussed in Sect. 6.4. The sudden tum-on of HI is so fast that the system has the same wave function or density matrix just after turn-on as it had before. The dipolar energy Ed which depends on the relative orientation of spins is thus the same at t = 0+ as at t = 0-. Moreover, since the dipolar Hamiltonian 1t~ is the same in both laboratory and rotating frame.
_ C1I2 H2 Ed = -~ = -Mo.::.L. 6] Ho
(6.68)
where 61 is the lattice temperature. The Zeeman energy is
Ez = -M·He = -Moho
(6.69) 237
The tolal energy. E = Ez + Ed, is then
E = -Mo{ho + HVHo) ~ - Moho
6.7 The Approach to Equilibrium for Weak HI (6.70)
A long lime later a spin temperature will be established together with a magnetizalion M parallel 10 He giving
E=_C(Hi;Hl) =_M(HJ;eHl )
(6.71)
But the lolal energy is conSClVed once HI has been turned on, so that, equating (6.70) and (6.71), we get
M = Mo H~O:~2
(6.72a)
L
o
Hillo
M z = NIo H2
H2
(6.72b)
M: = Mo H2 : H2
(6.72c)
0+
L
,,2 o
L
This equation shows that exactly at resonance, M would vanish. The null is vel)' sharp, M:t varying linearly with ho, so that observation of the null provides a precise method of observing exact resonance. II is interesting to contrast these results with conventional saturation theory. For a spin! system, M: is related to the population difference n by the equation /'fm
~-~~ 2 Saturation theory says that the equilibrium population difference, assuming infinite Tl' is n = O. hence M z = O. Equation (6.72c), on the other hand, says M z will be zero only when saturation is perfonned exactly at resonance (ho = 0). Conventional saturation theory assumes the transverse magnetization vanishes as well as M z . Equation (6.72b) shows that in general, M:I; is large, and may in fact be larger than M z . If (6.72) has a simple geomelTical meaning. After H] is pulsed on. M precesses about He. The component of M parallel to He cannot decay without energy exchange to the lattice. but Ihe componenl perpendicular can decay since Ihe local field gives a spread in precession frequencies. Thus, after several times l/-yHL' M will be parallel to lIe. and will have a magnitude given by the projeclion of the il'lilial Mo on He.
Hi:» Hl.
We saw in Sect. 6.5 that standard perturbation theory predicted that following the turn-on of a weak HI the population difference 11 would go to zero for long times, although we recognized that we could not rigorously apply perturbation theory to times greater than l/W. The requirement of a weak HI was necessary in order that perturbation theory be valid for at least short times. Of course, since M z is proportional to n. this implies M z would go to zero. In Sect. 6.6, however, we saw that M z would, under thesc conditions, go to an equilibrium valuc Mz)equil = Mo , 2
10
il 2 °H 2
+
(6.74)
L
fIr
where we have assumed Hr« 116 and «: HE (although we note that the equilibrium expressions in Sect. 6.6 were not limited to weak H J). It is therefore clear, as we suspected, that for fong limes, perturbation theory does flat give correct predictions. For short time intervals, however, it must be correct. Recalling the proportionality between M. and n, we know that
d~~,
_ -2W(w)M,
(6.75)
for times short compared to I/W(w). How can we describe M z for longer times? The solution to this problem was worked out by Provotorov [6.21] in an elegant paper utilizing powerful techniques. Rather than outlining his analysis, we will give an altemative derivation of his result. We note that in the absence of HI, thc Zeeman interaction in the rotating frame is just
1iz = -Iflhol z
(6.76)
Let us make the assumption that we can assign a tempcrature 0z to this Zeeman Hamiltonian, and 0d to the dipolar This assumption may not be rigorously correct, bUI it is al least simple, and corresponds to the facts at the time HJ is turned on. Immediatcly before turning on H] the dipolar system is at the lauice is the samc in thc rotating or laboratory frame. In temperature 01, since rhe laboratory frame we have
'Hl
1t3
CHo
Mo=-8,
(6.77)
But in the rotating frame
Clio Mo=-8z
(6.78)
Inasmuch as Ho ~ h o , Oz «0,. Thus the Zeeman temperature in the rotating frame, 0z, is very cold compared to the dipolar temperature tJ d . Turning on II] couples the two reservoirs and they approach the final equilibrium value given by the analysis of Sect. 6.6. The coupling of HI produces lTansitions between the
°
238
239
energy levels of the Zeeman and dipolar systems which we assume are governed by simple rate equations for the population of the various slates. (This assumption is similar to our postulates of Sect. 5.2. It is quite common in all cross-relaxation calculations. Provotorov makes it implicit in his work when he evaluates the relaxation times.) Since there are many states, a large number of coupled rate equations similar to (5.13) result. As has been shown by Schumacher [6.22], when the two systems are char· acterized by temperatures, the many equations reduce to twO coupled linear rate equations, one for (1/8z), the other for (I/8d) much as the many equations represented by (5.13) reduce to a single rate equation (5.27). But the conservation of energy gives a relationship between 8z and 8J :
Ch 2
CH2
8z
8d
___ 0 _ ~ ::::: const
(6.79)
Equation (6.79) is a first integral of the coupled equations, so that one of the resultant time constants is infinite, and a single exponential results. Since M, oc 118z, this means M~ relaxes according to a single exponential towards its equilibrium value. Using the fact that M, ::::: Mo initially, we get an equation for M, as a function of time,
M" - M,,)equil::::: (Mo - M~)equil)e-t/T
(6.80)
The only unknown in this equation is 1". We can, however, easily calculate il as follows. Taking the derivative of (6.80), evaluating it at t : : : 0, and comparing with (6.75) (which must be valid initially where perturbation theory is correcl), we get
~ T
Mo
:::::2W(w)
Ala
(6.81)
Mz)equil
Using (6.72c) for M')equil we get
2
2
2
_I ::::: 2W(w) h 0 +HL::::: 7(72 H2 (h 0 +H2) II 9(W) 1" H2 IH2 L
(6.82)
L
This result is the same as that firs! found by Provotorov, as indeed it must be since we have made the same approximations as he. The complete time development of the magnetization is therefore
M,::::: 2Mo 2 ho+HLexp -1r"f HI2(h~+Hf) 2 y(w)t )] ho+H L HL
[2
2 (
2
I
Slope b,OW" I,om f'tnuroutioll tlltory
". Filial rolue gi~fm by Uedfield O,eo,y
, Fig.6.6. M. "eUIl! t during saturation ~ given by the argument in ~he te,,~. The equlltion ror M. 115 Il runction or lime is round by joining lhe initilll region, where dM./dt is known rrom perturba~ion theory, to ~he equilibrium "alue, give" by Redfield theory, using Sc1wlllochu's observation lhat ~he approa<:h lO equilibrium involves a single exponential
6.8 Conditions for Validity of the Redfield Hypothesis We have noted that lhe concep! of spin temperalure in the rOlating frame has as a basic requirement neglect of certain lime-dependent !erms in the Hamil!onian
J
transformed to the rotating frame. This means, in essence H a,. Hi + Ht. On the basis of the previous seclion we can now add a second requirement. We saw that in the absence of HI> the equality of dipolar and Zeeman temperalures in the laboratory reference frame implied inequality in the rotating frame. Spinlattice relaxation will allempl 10 bring the two temperalUres together at the lanice temperature in laboratory frame in a rime TI. On the other hand, the presence of H I attempts to equalize them in rotating frame. The two tendencies are in conflict. The slrength of the tendency is indicated by the corresponding relaxation times (a short relaxation lime means a correspondingly strong tendency for the associated equilibrium). Thus a spin temperature will be established in the rotating frame only if 1" is much less Ihan Tt. Thus we find that a spin temperature is established in the rotating frame and that we muSt employ Redfield's approach provided
T, 1 -:» T
(6.83)
This expression is remarkable since it involves the successful integralion of the equations of motion well beyond the time (l/W) for which perturbation lheory is usually valid (see Fig. 6.6).
240
".
1r"'(2 Hr
or
(h~
;r
l )9(W)T1 ,. I
(6.84)
This is almost exactly the conventional condition for saturation. Note that the longer T I , the smaller the HI which will satisfy (6.84). In particular, frequently (6.84) is satisfied even when HI <. HI..
241
6.9 Spin-Lattice Effects So far we have considered the case of magnetic resonance on a time scale shon compared 10 the spin-lattice relaxation time. In many instances one performs
l
transient experiments which satisfy this condition. However, an equally important
I
case arises when one does experiments on a lime scale long compared (0 T. as when one employs steady-state apparatus. One can still salisfy the criterion for validity of the Redfield theory given by (6.84), but the spin temperature 8 in the rotating frame is now determined by the coupling to the lattice. As we shall see, this slatement by no means implies thai () = 8,. bUI only that it is a function of 8/_ Fortunmely it is very simple to generalize oue previous treatment 10 include the lattice. As a maHer of fact. in Redfield's famous paper he considered just this case, but we have put off consideration of the lattice relaxation until this Stage to simplify the discussion. In general, when the spin system exchanges energy with the lattice, the internal equilibrium of the spin systcm in the rotaling frame is momentarily disturbed. The basic assumption we make is that the cross·relax;lIion bctween the dipolar and Zeeman systems is so rapid compared to T I that, following an exchange of energy of the spins with the lattice, a new spin temperature is rapidly established. Thus, the lattice always finds the spin system described by a temperature in the rotating frame. We considcr that thcre are three basic relaxation equations for the classical magnetizations lI{zl and Mz: and for the expectation value of the dipolar energy (1t~), which we write down phenomenologically in a fonn chosen to assure that in the absence of HI the spin system would reach thennal equilibrim with the lattice.
·l
DMz:
7ft =
1110 -Mz: Tg
'Dt=
(1i?,), - (1f.\) Te
-E = -M,ho -
°
M.H, + (1id )
(6.88)
we can find the rate at which the lattice coupling changes E by taking Ihe time derivative of (6.88):
I
I
(6.85)
dE
DM~
a
0
(6.89)
1be only conttibutions to the time derivatives we need consider are those given by (6.85). (6.86), and (6.87) since other contributions do not change the tOlal energy E. Employing (6.85-87). assuming that M always lies along He' and using (6.58) and (6.59). we readily find an equation for the magnitude of M
dM
••
DMz:
. dt = -h0 /5t - H'/5t + at (1id )·
-
(6.86)
o(1i?,)
statistical mechanics for which there are a number of collision terms, each of which changes the distribution function f and for each one of which one could compute af/Dt. Thus. if the Zeeman energy -M· He and the dipolar energy (1t3) ..w ere not in thennal equilibrium with one another. the magnetization would grow (or shrink) unlil equilibrium was reached. ThaI process clearly produces a contribution to the three time derivatives (6.85-87) which we have not included in the equations. The lattice coupling. represented by Ta • T/>, and Te , must be expected to push the spin system out of thermal equilibrium in the rotating frame, but the internal couplings of the spin system counteract that trend by producing energy exchanges within the spin system. Changes within the spin system must conserve the total energy of the spins. If then we consider the rate of change of the expectation value of energy, E, we can ignore any changes which simply redistribute energy within the spin system. and keep only changes associated with the lattice. Using the fact that
dt
I = -(M", - M)
(6.9Oa)
Ttl
and for the spin temperature 8
(6.9Ob) (6.87)
where To and T/) and T e are relaxation times corresponding to exchange of energy with the lattice. and where (1t3)/ is the value of (1t3) when the spin temperature is equal to the lattice tcmperature. We have used partial derivative signs in these equations 10 emphasize that they represent the changes in these quantities induced by lauice coupling only. Thus. although Ta is the usual TI. T/) is lIot the usual T2 (::::: Il-yH t ,}. Tb is gcnerally of onler TI, a much longer time. A good analogy to these equations is to think of a Boltzmann equation in.
1 We do nol worry !Iboul Ai, since its relAxlllion does nOl c:hange lhe el1crgy in lhe rol.llling frame (in essenc:e we llSSume At, = 0).
where we have introduced a notation Tie and where
M
_ cq -
MoHerr{hoITa ) (hyTo ) + (H?ITb) + (HlJTc )
(6.910)
"
2 2 I (h H2 +:.:J. H ) _I = .::2. + _, Tl e h~+H?+Hl To Tb Tc
(6.9Ib)
[We have here neglected the term (1t~)1 of (6.87)). Note in particular that Mcq := 0 exactly at resonance, is poslUve when Ito";> 0 (i.e., M is parallel to He). but is negative for 110 < 0 (that is, M is antiparallel to He)' The last case corresponds to a negative spin temperature in I
242
243
the rotating frame. Equation (6.91a) shows that the equilibrium 8 is far from the lattice temperature 8/ and may even be of the opposite sign. Since Mo is inversely proponional 108/, it is still true Ihat 8f detennines 8. even Ihough they are quite different. The negative temperature one sometimes finds is a simple manifestation of the fact that M;: always tends to be positive whether ho is positive or negative. In fact. one can say that the equilibrium is reached as follows: The strong internal coupling of the spin system (which guarantees a spin temperature) keeps M along He, since Curie's law is a vector law. The lattice is anempting (a) to make Ihe z·component of M be Mo. but (b) the xcomponent be zero. (a) would make M bigger than MO so that its projection on the z-axis is Mo, whereas (b) would make M be zero. The lauice is thus fighting itself since (a) and (b) are inconsistent. The equilibrium value of (6.91a) results.
6.10 Spin Locking, T 1g, and Slow Motion As discussed above, and shown in Fig. 6.7, a spin temperature in the rotating frame is established in a rime T from an arbitrary initial condition wilhoUl exchange of energy with the lauice. But this is only a quasi-equilibrium value since. over the subsequent time TIL" the spin temperature changes as energy is exchanged with the lanice to drive M to M equil of (6.91 a) (see Fig. 6.7). During this process M lies along HeW' Thus if one starts at resonance having oriented M along HI, M will not decay in a time characterized by the inverse of the
M
line width, but rather with the time Tt, which requires energy exchange with the lattice. In principle. by going to low enough temperalUres one should be able to make T t , as long as one pleases. If one does this, magnetization along HeW in the..rotating frame. following at time T to establish a spin temperature, will remain without decay for as long as one has chosen to make TIL" This time may be seconds in metals or even hours in insulators. Even though we can lock the magnetization along Herr for a time T, when H] is on, if we remove H, suddenly, M wUl decay 10 zero ill a lime oi order of the inverse line width. That is, the Bloch equations with the usual meaning of T2 give a rough qualitative description of what happens. In contemplating the Redfield theory, it is helpfUl to go back to the SilUalion of the laboratory frame without alternating fields. There we see that smning with a magnetized system we can tum Ho to zero slowly and later tum it back up to its original value. When Ho = O. M = 0, but the full M is recovered when Ho is lUmed back on, all without exchange of energy with the lattice. In zero field the order is manifested by the preferential alignment of nuclear moments along the direction of the local fields of their neighbors. The ex.istence of order in the local fields is the basis of a technique [6.13] for observing motions which are much too slow to be seen as a T, minimum or a line·narrowing. Consider a nucleus #1 with a neighbor nucleus #2. Suppose now that #2 makes a sudden jump. as in the process of diffusion. The duration of the jump is perhaps 10- 12 to 10- 13 , very fast compared to nuclear precession frequencies. Thus the local field at #1 arising from #2 changes suddenly in both magnitude and direction. The orientation of the spin of nucleus #1 is thus somewhat randomized relative to the local field. If the mean time a given nucleus sits between jumps is T m , the alignment of nuclei in the local fields of neighbors, that is the ordered state, can persist only a time of T m . Thus to carry out a full demagnetiz..1tion and remagnetization cycle with full recovery of the initial magnetization, we must remain in zero field for a time less than T m. Of course even were there no jumping, any T] process would change the entropy of the spin system. so that in any event we must complete the cycle in a time less than T,. We can conclude i)
Vatu .. glvclI by (6.12a)
~
ii) Votu .. glv,," by (6.9Ia)
T~,1
Fig.6.7. The hierarehy of times followin& sudden tum-on of If l
• During time T, a spin temperature is established in the rolating frame, the value of which is determined by the expectation value of the ener&y, E. immediately after tUMIing on 11\. Over a longer time T 1I. the spin temperature in the rotatin& frame challSes, changing /If correspondingly in response to ener-gy transfer with the lattice. Both processes require Redfuftr s spin temperature hypothesis to analyze sina: we have M5umed T<:T'I
244
The demagnetization-remagnetization cycle can be used 10 monitor the zero field TI. We can detect jumping when T m < T •.
We can apply these same concepts to adiabatic demagnetization ex.periments perfonned in the rotating reference frame, as was done by Slichtcr and Ai/ion [6.23,24] and by Look and Lowe [6.25]. The analysis is quite straight forward. Since jumping causes sudden changes in the dipolar energy, it gives a contribution to IlTc which is
2
;;;--'--- = -(1 - p) Tc jumping T m
(6.92)
245
where p is a quantity expressing the fact that the local field after a jump is not completely random relative to its value before the jump. It can be calculated. Its value depends on the jump mechanism (vacancy, interstitial, etc.). The 2 arises because the local field is a coupling between pairs of nuclei, either of which can jump. Thus, simple measurements of T1(J enable one to measure a10mic motion or molecular reorientation when T m is comparable to T1(J arising from other causes.
7. Double Resonance
7.1 What Is Double Resonance and Why Do It? One of the most important developments beyond the original concept of magnetic resonance is so-called double resonance in which, as the name suggests, one excites one resonant transition of a system while simultaneously monitoring a different transition. There are many reasons for doing double resonance. The goals include polarizing nuclei, enhancing sensitivity, simplifying spectra, unravelling complex spectra, and generating coherent radiation (e.g. masers and lasers). The field has been characterized by a rich inventiveness which seems to continue unabated to the present day, and which makes the task of understanding all that has been done rather overwhelming. It is useful, therefore. to find some means of classifying the work into broad areas which involve related concepts. We shall employ three broad categories. The first category of double resonance makes use of spin-Iauice relaxation mechanisms. We call it the Pound-Overhauser double resonance after two of the important pioneers. The method involves a family of energy levels whose populations are ordinarily held in thennal equilibrium by thennal relaxation processes. If one saturates one of the transitions (1.34) one so-to-speak clamps their populations together (Le. forces them to be equal). The thennal relaxation processes then repopulate all the levels. producing unusual population differences which may possess useful properties (for example, the upper of two energy levels may have a larger population than the lower, or a population difference which is nonnally small may become large). Included in this category are methods of dynamic polarizaliotl of nuclei (the Overhauser effect and solid effec!), electronnuclear double resonance (ENDOR). and masers and lasers. A second category depends on cross-relaxation phenomena, hence we call it cross-relaxation double resonance. The fundamental concept is that if two spin systems can exchange energy (i.e. cross-relax), one can detect the absorption of energy by a resonant altemating field tllned to one spin system by lise of a second alternating field to monitor the temperature of a second spin system. Various experiments involving cycling the "static" magnetic fall in this category, as does the Hartmann-Hahn method which is the basis for the technique referred to today as "CP" (meaning "cross polarization") introduced by Pines, Gibby, and Waugh for enhancing the sensitivity of CI3 spectroscopy. [n general these techniques 246
247
require that some appropriate cross-relaxation process exist and that it be much faster than certain spin-Ianice relaxation processes which would otherwise destroy the effect. The third category depends in general on the existence of spin-spin couplings which in many cases must not be unduly obscured by either spin-lattice relaxation or cross-relaxation. We shall therefore call it spin coherence double resonance because it depends on the ability of spins to precess coherently for a sufficient time to reveal the spin-spin splittings. Typically, one here makes use of the faCI Ihat when two nuclei are coupled, changing the spin orientation of one nucleus changes the precession frequency of the nuclei to which it is coupled, so that the second nucleus can reveal in this way when Ihe first nucleus is being subjected to a resonant alternating magnetic field. Among the examples of this category are topics known as spin decoupling, spin tickling, spin echo double resonance (SEOOR), coherence transfer, and two-dimensional Fourier transform NMR (2D-Fr NMR). 1be last of these is fundamental 10 many important areas of resonance ranging from NMR imaging to delennining the struClure of complex biomolecules. Our purpose in introducing three classifications is pedagogical. Other scientists might pick different classifications. Of course, most schemes of classification are not perfect. For example, the nuclear Overhauser effect leads 10 effecls in 20FT NMR which are exceedingly useful and important. Moreover, 2D-Fr NMR involving only one nuclear species is not a double resonance experimenl (only one oscillator is used) but it can be conceptually viewed as one in which lhe ability of a large HI 10 excite all the nuclei in a spectrum obviates Ihe necessity of having a separale oscillator for each NMR line. It is not our intention in picking three classifications to imply that there is only one idea involved in each category. Indeed, within the groupings we employ there are numerous important innovations. For example, it was many ye~ after lhe firsl spin decoupling experiments were perfonned that 20 Fourier transfonn spectroscopy was invented.
7.2 Basic Elements of the Overhauser-Pound Family of Double Resonance The first double resonance experiment was carried out by Pound p.l] on the Na 23 nuclear resonance in NaN03. His aim was to prove that the spin lattice relax:ltion mechanism was via time-dependent electric field gradients. He computed the thermally induced transition probabilities between the various spin states based on this mechanism. In the NaN03 crystal, the Na resonance is split by an axially symmetrical electric field gradient into three distinct absorption lines corresponding to transitions between the m values of ~ to to and to He predicted and observed that saturating any Olle transition (e.g., the ~ to would produce level populations of all the states which differ from
-4 -l 4) 248
!' ! -4,
4
thermal equilibrium. Thus when he salurated Ihe j to transition, the intensity of the to transition increased by a factor of ~ times its normal intensity. We will discuss below why such intensity changes occur in connection with the Over.IJauser effeci. By means of a series of experiments saturating the various transitions while observing the OIhers, he verified that the spin-Iallice relaxation was via electric quadrupole coupling. The second double resonance experiment perfomled was by Carver (7.2). Ollerhauser, while still a graduate student at Berkeley, had predicted (7.3J that if one salurated the condUClion electron spin resonance in a metal, the nuclear spins would be polarized 10000foid more strongly than their normal polarization in the absence of electron saturation. (For an excellent summary of dynamic nuclear polarization by one of the important pioneers, see [7.4].) Crudely speaking, Overhauser predicted a polarization of the nuclei which they would have if the electron spin BollZmann factor were used in place of the nuclear spin Boltzmann factor. It may be hard for readers today to appreciate the deep scepticism with which Ollerha"ser' s proposal was greeted by the resonance community, because today there are many schemes for dynamic polarization, and the principle forms the basis for many other important techniques. However, in 1953 Ihere was a widespread (Ihough short-lived) belief Ihal Ollerhauser's scheme must violate the second law of thennodynamics. Carver's experiment was not only Ihe first demonstration of the dynamic polarization of nuclei, but perhaps more importantly showed that Ollerhauser's revolutionary Ihoughl was correct. His concept reoriented the thinking of resonators. While his concept stimulated the subsequent invention of other schemes of dynamic polarization, perhaps more imponant was the stimulus it provided to exploration of other novel effects of pumping transitions and doing double resonance. Figure 7.1 shows the first dynamic polarizalion of nuclei, the Li nuclei in lithium metal. To perform this experiment it is necessary to have an HI of 510 10 Gauss to salurate the condUClion electron spin resonance. One also wants a size of metal particle small compared to the skin depth at the electron frequency so Ihat aU the nuclei are polarized by the electron saturation. Accordingly, Carver used a solenoid 10 generate a static field of about 30Gauss, which put Ihe electron spin resonance at 84 MHz, and the lithium nuclear resonance at a frequency of 50kHz. At that, low frequency, the nuclear resonance was too weak to be observed directly (Fig.7.1a), but popped up from the noise instantly when the electron saturating oscillator was turned on.. Carver calibrated the degree of polarization by observing the proton resonance in mineral oil (Fig. 7.lc). He went on to show Ihal dynamic pol:uization did not require a melal by working with the proton resonance of liquid ammonia in which sodium had been dissolved. The sodium atoms in such solutions ionize, giving free electrons whose resonance is readily saturated. The principles underlying the Overhauser effect are in fact identical to those on which POUlld based his experiment. The Overhauser effect provided a strong impetus to the development of double resonance melhods, in part because Over-
4 -4
249
........."p.p.................
~
a
c
.
~:,
Fig.7.1. Demonstration of the Overhauscr nudear polarization effed on Li 7 nuclei in metallic lithium by Carver. The osdlloscope picture shows nuclear absorption plotted vertically versus magnetic field. The magnelic field excursion is about 0.2 Gauss. The top line shows the normal Li 7 nuclear resOnanCe (lost in noi~ at the 50kHz frequency of the NMR apparatus). The middle line shows the Li 7 nuclear resonance enhanced by saturating the electron spin resonance. The experimental conditions of (3) and (b) differ only in turning on the electron spin saturating oscillator. The bottom line shows the proton resonance from a glycerine sample containing eight times l'lS many protons under the same expt'rimenlal conditions, from which one concludes the J.i7 nuclear polarization was increased by a factor of 100
m,.
Hence we take the eigenvalues of I energy eigenvalues are then
E = "fefi.Homs
ms
=±~
z.as another good. quantum number. The
+Amlms - ",(n!lHom, m, = ±~
(7.3)
It is convenient to label states and wave functions by a convention
(7.4)
where c = 2mS, p. = 2m,. A state with ms = +~, m, = -~ is written as 1+ -). The selection rule for transitions induced by an applied alternating field is Llms = ± 1, Llm, = 0, or LlmS = 0, Llm, = ± 1. The first corresponds to an electron spin resonance, the second to nuclear resonance. (See Sect. 11.3 for a detailed discussion). Their resonance frequencies, We and Wn, are (7.5.)
hauser had made an ingenious and daring prediction which many talented physi-
(7.5b)
cists judged to be wrong, and in pan because he addressed a topic of interest to a community of scienlists much broader than resonators. Although Pound did not recognize that if his concept were applied to other systems it could lead to large nuclear polarizations, there can be lillie doubt that even without Overhauser's contribution, Pound's invention of double resonance would have caused a burgeoning of the field.
There are thus four allowed transitions. They are shown in Fig. 7.2. The relative size of IAI and ("fnhH) influences the appearance of the energy level diagram slightly as is shown in Fig. 7.3. If the nucleus and electron under consideration are far apart, typically IAI < l"fnfi.Hol (Fig.7.3a). rf the nucleus and electron are close together, usually IAI > hu rlHo I (Fig. 7.3b). The fonner case is encountered for a typical nucleus in a solid which has a low concentration of paramagnetic centers. The latter case is encountered for nuclei of paramagnetic atoms, or for nuclei which are very close neighbors of a pararragnetic center. For the rest of our discussion we adopt the convention of drawing figures which look like Fig. 7.2a, though we do not mean to imply thereby anything about the relative size of IAI and hn!lHol.
7.3 Energy Levels and Transitions of a Model System To understand the principles of the many applications, one can consider a very simple system consisting of a nucleus with spin 1= coupled to an electron of spin S = acted on by an external static magnetic field Ho. The Hamiltonian for this system is
!'
11. ='YehHoSz+AI.S-'YnhHo1z
!
where we have used subscripts e and n to denote electrons and nuclei, and where we have taken the form of electron-nuclear coupling appropriate for s-states. We assume that IcrlHo» A (the strong field approximation). Of course ,e» l"fnl. These assumptions make Sz nearly commute with 11.; hence mS, the eigenvalue of Sz, is a good. quantum number. Only the term AlzSz of the electron-nuclear coupling gives diagonal terms. so the Hamiltonian is effectively (7.2) 250
++1
(7.1)
-+~+i~1 (iI) Nudl'lIr r
I
w .. '" 'Y.. 11 0 +
.!!.I 211
->
(b) I;·/,.,.trlill V!ill "'SOllllllce 1rl1!l)'ilirJIIs
Fig.7.2a,b. The energy level diagram and allowed transitions of a system consisting of an electron of spin S = coupled to 3 nuclcus of spin I acted ou by an externa.t static magnetic field flo. The figure assumes 'Yn is negative
!
=t
251
++--
++--
The resul/ of using the more general form offunction of (7.8), whether as a result of solving the simpler HamUlonian of (7.1) more exac/ly, or as a result of solving the more general (7.7), is thai, on application of an applied alternating magnetic field, transi/iolls other t!lan those shown in Fig.7.2 become possible. We adopt the convention of calling transitions other than the four in Fig. 7.2 "forbidden transitions" .
--+-
--+-
-+--
In the absence of applied alternating fields, populations of the energy levels of the combined spin system are given by the Boltzmann factors when the system is in thennal equilibrium. As discussed in Chapter I, the achievement of thermal equilibrium can be thought of as resulting from lransitions induced by the coupling to a thennal reservoir in which the thermally induced transition probability We'l,c,y from state 1!"'1) to state Ie'l') is related to the rate Wc"'"e.,
-+-(a) DisrulI/
(h) N"urb)' 'IIICIt-,ts
IIlIdt'('$
Fig. 7 Ja b. The cITed of lhe size of the electron-nucle;o.f coupling II/h relative to the nucle'at ~nallce frequency Col on Lhe appearance of the energy level dillgn"us. For nuclei far from the dectron, lJ$ually iAI < h'.h 110 For nuclei ne~r. to the electron, .rrcquen~ly IAI > b.h/fol. The figu~ ll$IIumes a n~"tive 1'n and A positIVe. What would It look hke for positive "fa?
I·
by
~w~<~,~,<,'-'L'" = _P<'_._'
We have assumed a particularly simple Conn of electton spin-nuclear spin coupling, that which arises from the Fenni COnlact hyperfine expression. The most general Conn of spin-spin coupling would be obtained by adding also the dipolar coupling of Sect. 3.2 between the electron spin and the nuclear spin. Then, as is discussed in Chapter II, the Hamillonian would be
Wc'l'"e'l
where Pe'l is the thennal equilibrium probability of occupation of state le'l). Or, using the Boltzmann relation, We."c''l' =exp[(E ., - EC'l,)/k TJ c IV ,.,. "e.,
'H. = 'YellHOS~ + Az'z,Sz,Iz ' + AII'y,[IIS" + Azzdz'Sz' - 7n1lHo1n (7.6)
jmsm/)
(1.12)
e
where the axes Z', y', z' are a set of principal axes. Solution of the more general Hamiltonian of (7.6) still gives an energy level diagram which looks much like either Fig.7.2a or 7.2b as long as 'Ye1,Ho:> IAz'Z'I, IAy'y'I, and IAz'z.l. In this approximation. ms, the eigenvalue of Sz, is still a good number, but m" the eigenvalue of It, is not necessarily so. Thus the lowest order wave functions tPi (where i distinguishes the four states) may not be
tPi =
(1.11)
Pe'l
It becomes convenient at this point to switch to a more compact notation. Since there are only four states, we label them 1,2,3, or 4 as shown in Fig. 7.4. The notation lVij then corresponds to the thennally induced transition rate from state i to state j. Transitions Wij in which the electron spin is flipped but not the nuclear spin are shown in Fig. 7.5. As an aid to memory, transitions between I and 2 or between 3 and 4 are electron transitions, lransitions between I and 3 or 2 and 4 are nuclear transitions. Also, level 2 lies below level I, level 4 lies below level 3 in Fig. 7.4.
(1.1)
but may instead be linear combinations of such states.
tPi =
L
cimSnI/lmSm,)
(1.8)
1/1,=1++)
"'S,ml
If we keep the Hamiltonian (7.2) but solve it more exactly, we find that the states are of the fonn of (7.8) rather than of (7.7). Since there are still only four energy levels, the notation left) [i.e. 1+ +), I - +), 1+ -), 1- -)] still has validity since it suffices to distinguish four levels. However, while
1/t 1
'"
1-+)
I/t. '" 1--)
Fig. 7.4. Definition of the four states 1,2,3, and <\ in lerms of the earlier llotalioll le./)
(1.9)
in general Izleft) '" !Jllell) 252
(1.10)
,
253
Fig. 7.5. Some thermally induced mp' ping rates which invotve ~he transilion of the electron spin. The direction of the lIrrow indicates the direction of the trallsition corresponding to the rates 1V21 , 1V12 , ete.
(7. 13c)
(7.13<1)
Since the probabilities of occupation must add to one, we have that PI +P2+P3+P4 = I
'h=)--)
We seek a steady-state solution, so we set the left side of (7. I 3a-
7.4 The Overhauser Effect
PI
Though Overhauser' s original proposal pertained to polarization of nuclei in a metal. the principle can be illustrated by considering the model system discussed in the previous section. We take the simplest Hamiltonian, (7.1). with the solutions of (7.3). We assume that the principal relaxation mechanisms are those shown in Fig.7.6 which involve electron spin relaxation (WI 2, W21> W34, W 43) and a combined nucleus-electron spin flip (W23. W32) such as one obtains from the Fermi contact interaction in a metal as explained in Sect.5.3. An applied alter· nating field induces transitions of the electrons between levels I
(7.14)
Pt W12 + (pz - PI)We
(7.13.)
(7. 13b)
= 1>2
(7.15)
Equation (7.l3d) then gives us W
4J va =P4--
W"
which is the normal thennal equilibrium population ratio for this pair of states. Use of (7.I3d) in (7.l3c) gives W2J
P3=P2W32
(7.17)
which is the normal thennal equilibrium population of states 2 and 3. Thus for the family of transitions shown, while the saturation will change all the populations, it only affects the population ratio of the pair of states between which We acts. This result is a special property of the assumption that there are no W;/s which also couple stale 1 to states 3 or 4. Since the ratios PJ!P4 and P2/va are both thennal equilibrium, it follows that the ratio of p., to 1>2 is also thennal equilibrium. r"Or a pair of levels in thennal equilibrium Pj = I)ie(Ei-B;)/kT
::::: PiBij 1/1,-1+-)
(7.16)
(7.18.) (7.ISb)
which defines the Quantity Bij, the Boltznumn ratio (whence the symbol B). Note t.h~ convention on the order of the symbols E;, Ej of (7.18a) and the subscripts t,) on B;j. We therefore have PI = P2
(7.19)
Whence
Fig. 7.6. The Overhauscr effcet. Thermally induced transitions lVi; shown attempt to Illaintain thermal equilibrium. An applied al~ernaling field induces elcetron spin transitions at 8 ra~e IV., between the 8ta~es I and 2
254
(7.20)
255
We have stated that the Overhauser effect produces a nuclear polarization. Let us compute the nuclear polarization. In a state le11), the nuclear spin expectation value is '//2 (7]:< +1 or -t). Therefore. the average expectation value of nuclear spin I z • (Iz) is (7.21.)
(I,) ~ L::p;(iII,li)
This expression is the same as would be found for the nucleus if acted upon by the external field only.1 Comparing (7.23b) with (7.26), we see that electron saturation has increased (Iz)·.by the ratio
~ "
(I,) (lz)th,mn
"" ~(Pl +P2 - P3 - P4)
(7.21b)
l2-B23- B 24
(7.21c)
"" 2 2+B23+B24
To appreciatc the significance of this expression, we evaluate it in the high temperature approximation:
E-E I ) kT
1+
I ) ~ ~ (Eo - E,) + (E, - E,) ( z 2 4kT
A
2" + inhHo
so that
(E, -E1)+(Eo - E,) +(E, - E,) 4kT
Since E2 - E 1 "" -icliHo -
(Iz )lIICTlll =
256
lin hHo
2
2kT
Equation (7.28) is the result Overhauser originally predictcd for a metal. In a metal. one electron couplcs to many nuclei, so there is only a single electron spin resonance, not a resolved pair as in our example.
(7.25)
which, in the high temperature approximation, gives ~ ([z) therm - 2
(7.28)
lOne is at first surpriscd that the coupling to the dedron docs nol come into {/'\h <m since one can view lhe encrgy levels as though there were an effective magnetic field /fe/T acting on the nucleus givcn by
(7.23b) 2 4J.:T If the electron spin resonance were not being saturated, it is easy to show that the thermal equilibrium (Iz)
_
in
(7.24)
.!. iehHo
I B21 + 1 - Bn ~ B24
(Izhherm
The essential feature of the Overhauser polarization of nuclei is that there be a dominant nuolear relaxation process which requires a simultaneous nuclear spin flip and electron spin flip. In mctals, the strong role of the Fenni contact tcrm leads to relaxation processes in which the nuclear and the electron spins flip in opposite directions through terms such as I+ S- or [- S+. Even with a conventional dipole-dipole coupling there are terms [the tenns B, E, and F of (3.7)] which have this propeny of correlated flips if they can lead to relaxation. In liquids, the translational and molecular rotational degrees of freedom introduce a time dependence to the dipolar coupling which enables these terms to produce relaxation. Therefore, paramagnetic ions in liquids should lead to an Overhauser effect. As wc have remarked. Carver and Slichter [7.2] demonstrated this result using solutions of Na dissolved in liquid ammonia. The Na ions give up their valence electron, producing isolated electron spins which relax the H nuclei of
z
2 B 21 + I +B23+B2'1
ie
(7.22b)
Using the approximation that iehHo is much the largest term, we get
(Iz)therm ""
(Iz)
~'-'-~
7.5 The Overhauser Effect in Liquids: The Nuclear Overhauser Effect
(7.23.)
(I ) ~
which means the nucleus is polarized as though its magnetic moment were comparable to the much larger moment of an electron! If both electron transitions arc saturated simultaneously
(7.22.)
Now E3 - E2 "" iehHo +inhHo
E4 - E2 ""
(7.27)
2in
Jl eff = Jl o
A/2. this gives (7.26)
-
AlliS --
'Ynll
+t
-t,
Since IllS occurs with virtunlly e(111
257
the ammonia. A similar effect should arise with lWO nuclei. as was recognized by Nolcomb (7.5]. Bloch [7.6J, and Solomon [7.7], and was demonstrated by Solomon in a pioneering double resonance experiment in HF utilizing the HI and FI9 resonances. In Sect. 3.4 we saw that the dipolar coupling between nuclei in a solid could be used to obtain infonnalion about struclUre of molecules or solids. In liquids the dipolar coupling is averaged out owing to the rapid translational and rotational motion. Nevertheless. the dipolar coupling still plays an important role in detennining the relaxation times. It has tumed out that the nuclear Overhauser effeci (both stalic and transient) gives important information about which resonance lines conespond to nuclei that are physically close to one another in a molecule. It is thus playing an important role in struclUre detenninalions of complex molecules. especially when combined with two-dimensional Fourier transfoml methods. We therefore seek to understand the Overhauser effect in liquids in greater detail. In order to see more clearly how the liquid degrees of freedom make an Overhauser effect possible, we examine a concrete example, the nuclear Overhauser effect of a molecule containing two spin! nuclei. An example is HF, the system studied by Solomon. We shall demonstrate that saturating the resonance of one nucleus (the S species) produces a polarization of the Other species (the [ spins) and then show that if one disturbs one species from thennal equilibrium with a pulse, the effect is revealed in the transient behavior of the other species. It is, then. convenient to write the Hamiltonian as (7.29) where 'H 12(t) is the dipolar coupling between the two nuclei, augmented. perhaps, by any pseudoexchange or pseudodipolar coupling (Sect. 4.9). and where the explicit inclusion of t emphasizes that in liquids the dipolar coupling is in general time dependent since the radius vector from one nucleus to its neighbor within the molecule is continually changing direction. There is also coupling of the nuclei in one molecule with those in another. We omit these effects since we are concerned with demonstrating principles. Our calculation would therefore apply rigorously only to a case in which the molecules are present in low concentration in a liquid which does not otherwise contain nuclear spins. Since the Hamiltonian commutes with both It and St, we can label its eigenslates by mJ and ms, the eigenvalues of [z and St. The energy levels then become
E = -"YJ'iHomJ -,sliHoms + Amrms
(7.30)
where A is the pseudoexchange coefficient. the only tenn of 'H'12 which does not average to zero under the molecular tumbling. If A .,; O. the resonances of the I-nuclei or of the S-nuclei are doublets. We shall for simplicity assume that A is zero, SO that both the I resonance and the S resonance consist of single resonance lines. Figure 7.7 shows the energy levels. labeled by mrms, and also 258
4--
w w
FIg. 7.7. The energy levels ImsmJ> of II p"ir of nuclei (posilive .,'5). An IIpplied allernalins field induces lhe transition 1 to 2 and 3 to 4 in which rnJ remains fixed, but ms chanses. The lransition rllte IV ill defined in lhe text
l---.·
numbered 1. 2. 3. and 4. We assume that we are irradiating the S transition. producing a transition probability per second of W between states 1 and 2 and between 3 and 4. To find what this does 10 the [-spin polarization, we calculate (It) given by
,
(It) =
L
(il[:li)')i = ~(Pl + P2 - P3 - P4)
(7.3Ia)
.=1
In a similar way (S:) = !(PI + P3 - P2 - P4)
(7.3lb)
To find the Pi'S. we must solve the four rate equations for the populations Pi (i = I to 4)
dp,
dt = PI WH + P2 W 24
+ P3W34 - P4(W41 + W,n + W 43)
+
(7.32<1)
dp3
dt = PI W 13 + P2 W23 + 1'4 W43 - P3(W3t + W32 + W34)
+(P'1-P3)W
•
(7.32b)
etc. Adding these two equations. gives d(P3
+ p<j)
dt
= PI (W14 + W13) + P2(W24 + W23) - P3(W31 + W32) -J,,,(W
(7.33)
In the steady state (where all dp./dt's vanish). (7.32) shows that as W is increased. 1'3 and Pot become progressively more nearly equal. In a similar manner. PI and 1'2 become progressively more nearly equal. We therefore set PI = 1>2 and P3 = 1'4, obtaining from (7.33) ~=~
+ W13 + WZ4 + W23) F =1'1W.U + W31 + W"2 + W32 G
W14 (
(7.34)
259
which defines F and G. Now. utilizing (7.12) and (7.18)
To calculate these transition rates we need several equations. From (5.281) 1
(7.35)
and so on for W13. W2". and W23. Since EPi P" :::: PI (FIG), we get
= I.
PI
= P2.
P3
= p".
J
Gmk(T)eXp[ - i(m - k)TdT)
willi
-~
and
(7.300)
hence
+~
Wkm = 2" _ h
Gm.(T) " (ml1i l (l)!k)(kl1i , (t + T)lm)
Taking Gmk(T) = Gmk(O)eXP(-jTI/Td
(7.36b)
PI = 2(1 + FIG)
W
Therefore
km
I (I-(FIG») PI - Poi = '2 1 + (FIG)
(7.37)
Utilizing PI = 1>2 and])3 = P'I> (7.31) becomes (7.38)
Utilizing (7.34) and (7.35) we can get an exact expression for F and G. However, in the case of a liquid. the high temperature approximation is valid even for electrons (perhaps liquid He is an exception!) so that Bij = I + (Ej - Ej)/kT
giving =
_I_{ 2kT
(7.39)
(7.40)
1- -) to
1+ -)
(7.41)
As we shall see, W42 = W31, so that .
(7.43)
and recalling that
B = -t(S- I+ + S+ r)(1 - 3 cos 2 8)
E = -i(S+ J+) sin 2
ee- 2i6
~~ (/6) ((1 -
3 cos' 8)'),.(+ -1[+5-1 -
+X- + Ir5+1 + -) (7.44)
to 1+ +) in which the I-spin flips up, W 42 is the transition from 1- -) to 1- +). also a transition in which the I*spin flips up. Finally W32 is the transition from 1+ -) to 1 - +) in which the S-spin flips down and the I-spin flips up. We shall define a new notation which makes these processes more explicit by defining U/l1 where M is the totnl chnnge in mr + ms in the tmnsition:
Ul(W) = W31(W)
k)2 T ;
For dipolar coupling, the 'H.1(t)'S represent the various tenns A. B, C, etc. from (3.7). Defining
Uo =
Examination of Fig.7.7 shows that W41 is llie transition from
+) in which both spins flip up. W31 is the uansition from
2Tc
+ (m
we get that
E
W"I(E4 I ) + W31(EJ - E I ) 2(W41 + W31 + W 42 + W32)
+ W 42(£4 - £2) + W32(£3 - E2)} 2(W"1 + W31 + W 4 2 + W 3 2)
1+
= Gmk(O) h2 I
Ao = "'IJ1sh 2/r 3
IG-F (I,) = (P, - p,) = '2 G + F
(I,)
we get from (5.297)
where «(1 - 3 cos 2 8)2)" .. means the average over the 411" solid angle of (I 3 cos 2 8)2. Thus «1_3cos2 8)2)"
•
=~J(I-3COS28)2sin8d8d¢=4/5 4.
(7.45)
Likewise (sin 2 8 cos 2 (J),,'/I" = 2/15
(7.46)
(sin" (J)o\?r:::: 8/15. In this manner we get Uo =
A~ -.!....
U2 =
h2 5"
(7.42)
Tc
11.2 10 1 +(ws -wr)2rt
A33
Tc
I +(wr +wS)2 Tt
(7.47) 261
The resulting polarization of the I-spins is thus, using (7.40), (7.41), and (7.47),
Now, in thennal equilibrium we have from Pj(T) '" PiO')B ij thai (7.55)
Pj(T) - ]>;(1') = Pi(T)(Bij - I) = Pi(Tkij
(I,)
Therefore, in (7.54) we recognize that since Pi (7.48)
';:!
Pi(T) S!'
t
W 4 1P\e14 S!' W 4 1[P4(T) - Pl(T>]
(7.56)
Substituting into (7.54) we gel Suppose
Iwsl »wl (as for S being an electron spin, in which case ws is negative)
and thai W~7";
«: I. Then
(I ) = ''"'5 (3/5 - 1/10) = ~ "W5 z 41.:1' (3/5 + 1110) 7 41.:1'
(7.50)
which for negative Ws is posltlve. Therefore, dipolar relaxation produces an Overhauser effect of the opposite sign to the conventional Overhauser effect since for dipolar coupling U2 »Uo. We have carried out the solution for the steady-state polarization of the 1· spins. If one is doing pulsed experiments, as is frequently the case for double resonance experiments or for two-dimensional Fourier transform experiments, the pulses disturb the 1-S spin system from thermal equilibrium, following which the thennal processes bring the system back to thermal equilibrium. One then speaks of a transient nuclear Overhauser effect. Thus, in general, one wishes to find the time dependence of the observables (I~(t) or (S~(t) for some sort of initial conditions. Solomon, in [7.7J, calculates the transient response and demonstrates what happens in a set of classic experiments. To do the calculation, one needs to solve for the time dependence of the Pi'S, starting with equations such as (7.32) with W = O. Thus we have
dp4
dt =PJWI4 +P2 W 24 +]J3 W 34 -P'I(W41 + W"2+ W"3)
(7.51)
- Pol -
[PI (T) - 1'4(T)j}
+ W",{)" - P4 - [p,(T) - p4(T)j} + W",{p3 -Pol - U'3(T) - 1',1 (T)]}
(7.49)
which is negative for electrons. On the other hand, if there were a strong Fenni contact tenn, so that Uo »UI or U2, as with the conventional Overhauscr effect,
( I ) = _ lIws z 4k1'
dP4 dt = W ll {PI
with similar equations for dpddt for i = I, 2, and 3. We saw in Chap. I that the fact that Wij oF Wj; is important in producing a thermal equilibrium population. That is the physical significance of (7.56). Clearly the family of equations represented by (7.54) describes a system which will relax to the thermal equilibrium populations p/T). We can then replace the Wij'S by the Um's, taking
Uo
W32 s:' W23 =
W42 s:' W2.1 = VI = W3l S! W I 3 W31 s:' W l 3 = U2
and introducing
Ui
(7.58a)
by Jhe relations
WI2 s:' W21 ==
ui = W34
s:' W"3
EI
cl" =
where
~E4
kT
relationships, one can now show Ihal d dt (Pl +!J2 -1'3 -1',,) = 2UI [(pol -]>2) + (P3 - PI)]
+ 2UO(p3 - 1") + 2U,(P4 - P'> - 2U I [PI (T) - P2(1') + P3(T) - PI (T)] - 2Uo[P,(T) - p,(T)]- 2U I [P,,(T) - p,(T)]
(7.52)
From (7.31) one can show that
Therefore (7.59) can be rewritten utilizing 10 and So to represent the thermal equilibrium values of (I~) and (S~) as
(7.53) = [10 _ (1,)](U0 + 2UI + U,) + [So - (5,)](U, - Uo) dt In a similar manner, we get
d(I,)
dp4 dt = W 41 (PI - P4) + W 42U>2 - p,,) + W 43 (1'3 - P4)
,,,
(7.59)
(7,60)
we get
+ W41PIC14 + W42P2C24 + W43P3C34
(7.58b)
U: differs from UI (7.47) by the subslitution of Ws for WI. Making use of these
Writing Wu = W<\lB I4 and taking BI4 = I +cH
(7.57)
(7,54)
(7,61,)
(7.61b)
'63
These two coupled linear differential equations have solutions consisling of real exponentials. They may be viewed as a nonnal modes problem with imaginary frequencies, i.e. real exponentials. There are thus two time constants. These equations show that as long as (5,) = 50 and (I,) = 10, both d(5,)/dt and d(I,)/dt remain zero. However, if one disturbs either spin system from thennal equilibrium, both spin systems will respond. Thus, if one makes ($,:) = 0 by applying a 7fn pulse to the 5-spins, not only will d(S,)/dt be differem from zero, but also d(l,)/dt will be nonzero. In general, then, when one applies pulses to a system of coupled spins, all spin populations will be found to respond as a result of the relaxation. We pose the detailed solutions of these rate equations as a homework problem.
~,
=1++)
I,
,, , WI':, ,,, , ~1=1-+)
Woo
,,, , ,,, IV4.J: : 11'.,. ,, ,, ,, t
, ,,, w,. ,,
~J=t+-)
~.=r--)
Fig. 7.3. The use or • rorbidden tnonsilion which flips oolh lile dN:tron and nucleus simultaneously, !Vea , to produce nuclear polilfiZlltioll. h is I181SUl'll«I thallransilions which ilwolve an e1edfOn sl>in-nil> only are lhe only sip;nilicallt lhermlll PI"OCes6e$
7.6 Polarization by Forbidden Transitions: The Solid Effect In order for the Overhauser effect to work, the nuclear relaxation process (such as W13, W24) cannot be allowed to short circuit relaxation in which both an electron and a nucleus flip, such as W23 or W!1. It is not always possible to meet those conditions. The condition which one can, in general, be quite sure will hold is that pure electron spin-flip processes (such as W 12 or W34) are much the fastest W;j's because electrons couple more strongly to tile lanice than do nuclei. Jeffries [7.8] and independently Abragam et al. (7.9) recognized that the so-called forbidden transitions were not strictly forbidden in many useful cases, and that one could use them to good advantage in achieving polarization. In fact, Erb et al. (7.10) independently discovered the effect experimemally. Using this scheme, Jeffries and his colleagues at Berkeley, and experimentalists at Saclay collaborating with Abragam. have obtained proton polarizations of over 70%, and have made a rich variety of applications. Invention of this technique was another major step forward. The phenomenon is often referred to as the solid effect. There are two possible forbidden transitions. They are shown in Figs. 7.8 and 7.9. The transition of Fig. 7.8 can be induced by an alternating field parallel to the static field when the simple isotrOpic electron-nuclear coupling of 0.1) is solved to the next higher order to include the effect of the tenn AUz 5 z + IySy ) in admixing 1+ -) with 1- +). The transitions of both Figs. 7.8 and 7.9 can be induced by alternating fields perpendicular to ]fo when the more general Hamiltonian, (7.6), is solved to adequate precision. For example, the dipole-dipole coupling between a nucleus a distrance r from the electron makes the transition matrix elements of Fig. 7.8 and 7.9 of order 'Yefilr 3 Ho times the matrix element of Fig.7.6 in which only an electron is flipped. The ratio also depends on the angle the static field makes with the axis connecting the nucleus and the electron. (The effect of the dipolar coupling is expressible more precisely in terms of the contribution of the dipoledipole coupling components AZ'Zl, Ay'y" and A,,'z')' 264
FIg. 7.'). A rorbidden transition 1Iltemalive to that or Fig. 7.8, which producl'3 Iloclear polarization or the opposite sign
Even though the transition probability Wen may be small, it is frequently possible to achieve strong enough alternating magnetic fields to make it larger than the thennal transition rates Wij in which a nucleus flips, so that the transition connected by Wen produces effective population equalization. Lei us analyze the case of Fig. 7.9 for which the transition between states I (-IPt = 1+ +)] and 4(IP4 = 1- -)) is saturated. Since we assume Ihe only thermal transitions of consequence are those shown, we immediately can write down 1'1 = P4
(7.62,) (7.62b) (7.62c) (7.63,)
(7.63b)
265
P3 :
~~B::,4",-J~_
(7.630)
2+8 12+ B 43
relative to the donor atom. The solution is a straightforward generaljzation of that of Sect. 7.3, using the N quantum numbers mi, the eigenvalues of fZi' ..E = '1ehHOmS +
Using (7.2Ib),
) 1 812 - B
(7 64) .
To appreciate the meaning of this expression we look .at l~e high temperature limit (evaluation of the expression for all temperatures IS given as a homework
Electron resonance occurs when Ams ""
± I, Ami"" 0 for all i. 1'he frequency
We
= "feHo
A;
+ ~ -;;m j
•
I 7enHO
2
(765)
2kT
.
so that the enhancement over the normal polarization is (7.66)
the full Overhauser effect.
7.7 Electron-Nuclear Double Resonance (ENDOR) A double resonance experiment of great historical importance was performed by George Feher [7.11]. He named it electron-nuclear double resonance (ENDOR). The purpose of this technique is to resolve otherwise unresolvable resonance lines. The concept involves in essence observing a nuclear resonance through its effect on an electron spin resonance. Fehu was studying the electron spin resonance of electrons bound to donor atoms in silicon. 5 % of the silicon nuclei, the isotope Si 29 with spin ~, possess a nuclear magnetic mome~1. As a result of the large radius of the orbit, the electron spin couples magneucally to many Si29 nuclei. Since the Si 29 nucleus can have two spin orientations, and since the Si29 hyperfine fields depend on where in the donor orbit the nucleus sits, there are many different local fields Si 29 nuclei can produce at the electron. The essential situation is shown by considering a Hamiltonian of the one electron interacting with N Si 29 spins. (For simplicity, to avoid the needed discussion of the statistical effect of an isotopic abundance, let us suppose all Si sites had Si 29 atoms.) We take it to be of the foml N
11. "" "fchHoSz + S·
L
A;Ii - "fnhHolz
(7.67)
i=1
where Ai is the hyperfine coupling of the ith nucleus (for simplicity we assume a simple Fermi contact form of coupling rather than the more general tensor fonn). The Ai'S are determined by the crystallographic location of the ith nucleus
(7.68)
We IS
problem).
(1,) :
I: AjmjmS - I: "fnhHOmj
(7.69)
+!
Since there are many possible values of Ai, and since each mi can be or (7.69) describes many frequencies, in fact virtually a continuum, rather than a set of resolved lines such as in the case of an electron interacting with only a single nucleus. For a sample consisting of many donors, there are 2 N various ways of assigning the mj's. The problem of resolving the hyperfine lines may be likened to that of a man with several telephones on his desk, all of which ring at the same time. If he tries to answer them all, he hears a jumble of conversations as all the callers speak to him at once. Of course his callers have no problem - they hear only one voice, though he hears several. Feher recognized lhal each Si29 nucleus experiences the hyperfine field of only one electron. Thus the nuclear resonances are sharp. Each Si 29 site gives rise to two resonance transitions, corresponding to whether the electron hyperfine field aids or opposes the applied magnetic field:
-i,
Wnj ""
"fnHo - AjmS
(7.70)
The nuclear transition frequencies of Si 29 nuclei located at each distinct crystallographic location relative to the donor atom give rise to a distinct pair of lines. For many sites, the Aj is sufficiently large that these lines are well resolved from the other transitions corresponding to other sites. Recognition that nuclear resonance would give resolved lines even though electron spin resonance did not was a great insight on Feher's part. But, however, the nuclear resonance signal would be weak since the donor concentration (_10 17 per cm 3) is typically so low. Feher conceived of the clever method of detecting the nuclear resonance 1;ly its effect on the electron spin resonance. Consider the situation in Fig. 7.10. We show 4 levels only, which represent the four energies we get if we hold all mi's fixed except for those of the jth nucleus:
L
E '" CrehHo +
Ajmj)ms + Apnpns - "fn1iJIOmj
(7.71)
i • j
Clearly we can define an effective static field acting on the electron of Heff = Ho +
A-kmj ji-j"fe
L
(7.72)
for this case. 266
267
'It,-I++l
Tn, , :,,'/', ill , '
..~ I
.,1' I... I ' I I
'¥l-l-+)
7.8 Bloembergen's Three-Level Maser
.;...
,"" ",""
,
, I
,I· I',
I"
I I t
I" I
'
p,
p,
'v,
'e)
..Ll..
(,)
1J>l '" (x/!
E' -E') (-;;:r-
J'1g. 7.10a-c. Ft~,'s scheme for observing nuclear rcsomlll(e by nlOniloring il.$ effcct on the electron spin resonance. flg\lrc 7.IOa shows lhe stales of (7.71) for the electron and the jlh nucleus, with the eleclron transition being shown between stales I llnd 2, and thermal relaxation pro«sscs such as in the Overhauser efrect. Figure 7.IOb shows the populations produced by saturating the electron transition. Figure 7.1Oe: shoWlii the effcct of an Ildiah2.tie pIlIJ!:IlI.ge through the resonant frequency of states 1 and 3, with lV~ turned off. When 'V~ is turned back on, it finds the electron transition bct"'een stales I and 2 momentarily ul1!laturated
We assume for concreteness that we have the same thennal relaxalion processes as with the Overhauser effect (Sect. 7.4), and saturate the same transition, The resulting populations are shown in Fig.7.10b. Suppose we momentarily switched off We' switched on the nuclear oscillator (Wn ) close to the transition frequency between states 1 and 3, and did an adiabatic passage with W n through the nuclear resonance line. The passage will exchange the populations between states 1 and 3, producing the populations shown in Fig. 7.IOc. If one then turns on We, initially the electron resonance is no longer fully saturated, so the elec· tron spin resonance has a changed signal height which decays as the transition saturates. Thus, the nuclear resonance produces an effect on the electron spin resonance. We have described a panicular method of doing ENDOR. There are many variations. The technique has been exceedingly important in mapping wave functions of a variety of paramagnetic centers, and thus is one of the principal ways one knows the structure of many important point imperfections in solids. (For an excellent review of the use of ESR and ENDOR to study structure of point imperfections, see [7.12].) It is of great importance for structural detenninations of biological molecules. For his discovery of ENDOR and applications, Feher was awarded the Buckley Prize in Solid State Physics of the American Physical Society.
A maser (microwave amplification by stimulated emission of radiation) works on me principle that irradiation of an absorption line associated with transilion between two energy levels will lead to a net emission of energy when the upper ~evel is more highly populated than the lower one. Townes (7.13]. and [7.14J, and Independently Prokhorov and Baso'll [7.15J, recognized that if such a situation could be achieved, it could be used as the basis of a new type of oscillmor. Townes achieved population inversion with ammonia gas by physically separating lhe molecules in the upper energy level from Ihe lower energy level by means of properly shaped electric fields, in essence a molecular beam apparatus. The frequency of the oscillator fell in the microwave region. Shonly Ihereafler, Bloembergen recognized Ihat one could produce the energy level inversion by use of the Overhauser-Pound principle of level repopulation through saturation. It required a system with three or more energy levels (Fig. 7.11).
,, , ,,
,pV
,, ,,] 11/ ,
J1
,J hIll
21
FIg. 7.1t. Bl«mbugtfl'll three-level maser. The thermal relaxation is shown by the dashed arrows, the induced transition by the twoheaded solid arrow. If the populations PI and PJ arc cquali:.:cd by saluration, Pl reaches a value which ClI.LJses lin inverted population either of 1)1 rell\live to P' or of p, relative to p:!, depending on lhe relative strenglh of lhermal rates bclwe<:n stales I lLnd 2 versus 2 and 3
If, for simplicity, we can neglect the thennal relaxation between states 2 and 3, then (7.73a)
PI = P3 P2=P I B I 2
PI = P' = P2=
so that 1
:;-;-''n2+ B12
Du 2+B 12
(7.73b) (7.74.)
(7.74b)
so that P2/P3 = B 12. Since
B I2 = e(E,-E,)/kT > 1 268
(7.75) 269
we have that
">1 P3
a~
making the upper level more highly populated than the lower, the condition for maser operation at the frequency (E 1 - ~)lf1. B/oembergen pointed out that the theoretical situation he described could be achieved with paramagnetic ions, as was shonly demonstrated [7.16]. Bfoembergetl's concept has been of enonnous value. Level inversion by pumping another transilion is the principle on which all lasers operate. The pumping may be done optically, as when a light source is used, or by the equivalent of pumping (which generates nonequilibrium populations between a pair of levels) in chemical reactions (chemical lasers) or electric discharges (electric discharge laser). Prokhorov and Basov [7.15] independently developed many of the same ideas.
7.9 The Problem of Sensitivity Sooner or later all resonators want to observe a resonance wh.ich is too weak to be seen. Resonances may be weak because the number of spins is low, for example since the isotopic species is rare, or the nuclei occupy special positions such as being surface atoms on a crystal or neighbors of an imperfection. The resonances may be weak because the nuclear ""( is small. The development of superconducting solenoids with their high magnetic fields has helped enormously in increasing sensitivity of NMR equipment because the quanta absorbed in resonance are thereby increased. Reduction of noise by narrow banding also helps. The practical problem used to be that a narrow bandwidth was achieved by using long integrating time constants, and those in tum meant it took a long time to sweep across the resonance line. By the time one was near the end of the sweep, the apparatus was likely to have drifted, giving distorted line shapes or spurious signals from drift in the apparatus baseline. The advent of the multichannel digital signal averager has dramatically changed all that by permitting cumulative addition of a number of rapid sweeps. If the appamtus drifts, one stops sweeping, readjusts the equipment, and continues sweeping. Moreover one sweeps only as long as is needed to achicve the desired signal-to-noise ratio. TIle signal accumulated is proportional to the num~ ber of sweeps N. Since the noise is random and both positive and negative, its amplitude increascs as .JFi, giving an overall increase in signal-to-noise ratio proportional to NI.JFi = .JFi. As a practical maller, this means one must increase the averaging time fourfold to gain a mere factor of 2 in signal-to-noise. This increase may not matter if it means going from I min to 4 min, but if one is already averaging for 270
I h, going to 4 h is a hanl way to improve signal·to-noise! These facts illustrate that when one is using averaging there is a big premium on setting variables such as HI> modulation amplitude, sweep excursion, etc., to the optimum. A corollary is when looking for an unknown, be willing to set the equipment to give ~ maximum signal even though you may thereby distort the resonance line (for example, from using too large a modulation, or from panially saturating). After the resonance has becn found you can focus attention on it with parameters adjusted to avoid distonion.
7.10 Cross-Relaxation Double Resonance No mailer what one does to improve one's apparatus, one eventually reaches the limit of the current state of experimental an. What then if the signal is still too weak to see? We tum now to the use of double resonance, assuming there are two resonances which can be excited, one the "weak" one too difficult to see directly, the other a "strong" one observable by conventional means. For example, the weak one might result from a low abundance species with spin S, whereas the strong one might result from an abundant species with spin I. (For simplicity, we use the term "rare" to imply the species whose resonance is weak. Since a low ""( could also make the resonance weak, "low ""(" can be substituted for "rare" in most cases.) If an H I is applied at resonance to the rare species, those nuclei absorb energy and their spin temperature rises. If the rare and abundant spins could exchange energy, the abundant spins would thereby get holter, and their signal amplitude would diminish. If the abundant species is thermally isolated from the outside world (i.e., has a long TI), this temperature rise can be made quite large and thus readily observable merely by silling on the weak resonance long enough. By going to low temperatures, TI of many systems can be made exceedingly long. Thus the crucial question becomes how well can the two species exchange energy. The mixing of the Iwo spin systems was sludied by Abragam and Proctor [7.17]. They employed the Li 7 and FI9 resonances in LiF. Their experiments were part of their studies of the fundamentals of spin temperature and of adiabatic demagnetization. Working in a field of several thousand gauss to observe the resonances, they prepared the system in some nonequilibrium state in the strong field (for example, they inverted the F I9 magnctiz..'ltion), removed the sample to a lower static field, allowed the Lj7 and FI9 spins to mix, then returned the sample to the strong field for inspection of the Li 7 and F19 resonances. In this way they found that in a field of 75 Gauss the spins come to a common spin temperature in a mixing time of 6s, that the time was unobservably shon at 30Gauss, and longer than the TI 's (several minutes) above lOOGauss.
271
A detailed study of cross-relaxation was made by Bloembergen et al. [7.18] and by Pershall [7.19] for both electron spin systems and nuclear spin systems. They show that the crucial problem is the failure of the Zeeman energy to match when two different nuclear species undergo mutual spin Aips. The mismatch in Zeeman energy must be made up by the dipolar coupling between the spins. The simplest sort of process arises when two nuclei have nearly the same 'Y. For example, HI and F I9 have 'Y's which differ by 5%. Owing to the existence of tenns such as S+ I- or S- 1+ in the dipolar coupling [the B tenn of the dipolar Hamiltonian of (3.7)], the dipolar coupling couples states in which the proton spin flips up (down) and the fluorine spin flips down (up). If we could consider the 'Y's as being identical, we would then have a situation such as shown in Fig. 7.12. Because the individual energy levels match, the initial state of the two spins, indicated by the two x's, has the same energy as the final state indicated by the two o's. A system started in the x state will undergo a transition to the o state by means of the S+ I- part of the B term in the dipolar coupling. For HI and F t9 , the fact that the 'Y's differ by 5 % means that in a strong magnetic field such as 10 kGauss, the Zeeman energy difference between the x and the o states would correspond to the energy of FI9 or HI in a field of 500Gauss. Such an energy mismatch would prevent the transition unless there were some other energy reservoir whose energy could change to make up the difference. A possible candidate is the dipolar energy reservoir. For a pair of spins, it has a typical value of 'Yl'YSh2/r3, which, expressed in units of magnetic field, is only a few gauss for reasonable values of r. Thus, it cannot make up for an energy mismatch of 500Gauss. If, however, the static field were much lower, the mismatch would be correspondingly reduced, and the mutual flips might become possible! Returning to LiF in a field of IOkGauss, the Li 7 resonance occurs at 16.547 MHz and the Ft9 resonance occurs at 40.055 MHz, a ratio of 2.420. Clearly the mismatch here is even worse than for H t and F t 9. What Bloembergen and his colleagues recognized [7.18,19] was thai 2.4 is close to 2.0, hence a process in which two U 7 nuclei flip up and one F I9 nucleus flips down comes much closer to satisfying energy conservation than a process in which only one Li7 flips. Now, Li 7 has a spin of ~' but for the sake of argument we are going to pretend it has a spin since the explanation is then simpler. To estimate the rate, one must utilize a fonnula for transition probabilities such as (2.182) (with
!'
Ms
=
M l = -1/2
-1/2
M s = +1/2 Spccies 5
272
Specic" J
]<'ig.7.12. A mulual spin-IIiI' from thc (x) lo lhe (D) slale cxchangcs energy belween the [-spins and the S-spins. Shown verlically are lhe allowed energy levels. For lhis cxample thc cnergy spacing is assumed to l>c lhe same for lhe two spc<.:ies, and bolh nudei have spin
t
0). One therefore looks for a matrix element which connects the initial state (two Li 7 spins up, one FI9 spin down) to the final state (two Li7 spins down, one FI9 spin up). If we denote the Li 7 by S and the FI9 by I, and use subscripts i, j, k to distinguish individual nuclei, a matrix element of the dipolar !enn B. (aIS; Ib), between states a and b flips only olle U 7 spin. To describe two spin flips we must go to second-order perturbation to get the effective matrix element (al?ierrlb) joining the initial and final states:
W =
t:
(al1iofflb) =
L: (al1i~I'~'~1idlb) °
b
°
where one matrix element might involve
st I":
(7.77)
st IZj and the other st Ii. There
are other possibilities, for example combined with SziSt The energy difference Eb - E c depends on what the initial and intennediate states are, but is of the order of magnitude of the Zeeman energies involved. In general, if one considers a process Wilh I1.F fluorine spin flips down and n Li Li spin flips up, with !lwLi or flWF energy change per Li or per F spin, the energy mismatch L!E between initial and final states is (7.78) (7.79)
nLi = 2
dE nwF =0.17
but for llr =
2
Hl.i =
5
11E = 0.066 flWF
Thus, the more spins we allow to flip simultaneously, the closer we can make the energies of initial and final states match. However, more spins flipping requires more dipolar terms acting simultaneously and thus requires third- or higher-order perturbation expressions. The first-order matrix elements are of order HI., where HI. is some sort of dipolar field of one nucleus at a neighbor. The secondorder expression (7.77) is then of order Hl/Ho, where Ho is the static field. Each increase in order of perturbation contributes a factor HLlHo. Thus, since H L
.'ig.7.0. Dah. or Pershiln for the cr0S5relaxation Lime 121 ror l,i7_1~19 in Li(~ versus static magnetic field for three crystal orientations (sec 17.191)
T1\ YS. magnetic rield in LiF
l /
10
When one has a single crystal in which some nuclei have a quadrupole splilling and others do not, there are magnetic fields other than zero in which the energy level splittings of the two systems become equaL Cross-relaxation is then r.apid. Edmonds {7.26] has reviewed nuclear quadrupole double resonance.
/ J 7.11 The Bloembergen-Sorokin Experiment
••< B ••
l,,
I
0.1
/ ,t
,I /
o.Oll-~-l('-7:-~--~ o
SO
Ma9"etil: field IGallssl
'"
Clearly the use of cross-relaxation as part of double resonance leads in a natural way to cycling the static field between low values for cross-relaxation and high values for observation. Anderson [7.20). working with Redfield, combined field cycling with Ihe application of an audio-frequency magnetic field applied while the static field was zero to heal the spins to plot out the zero-field absorption characteristics of a spin system. Thus they used field cycling to give them the sensitivity of resonance in a strong field to monitor the effects they produced in zero field. Redfield [7.21], Fernelius [7.22], Slusher and Hahn [7.23], Minier [7.24], and others utilized field cycling to observe quadrupole splittings of nuclei near to foreign atoms. If one has a sample which is a powder (e.g. a metal sample where one employs a powder to overcome problems of the skin effect), an advantage of demagnetizing to zero field is that the quadrupole splittings are then identical in all the individual crystallites. There is therefore no longer any powder broadening of the NMR transitions [7.21-24]. To observe very large quadrupole splittings in powders, there is a problem of coupling the energy absorbed by the quadruple transition into the rest of the spin system. It is necessary then to adjust the amplitude of the "audio" frequency alternating magnetic field according to principles of Hahn and Hartmann discussed later in this chapter. Weitekamp et al. [7.25} have employed field cycling to zero field together with Fourier transfom NMR to simplify dipolar spectra of powdered samples. 274
An important advance in cross-relaxation double resonance was discovered by Bloembergen aod Sorokin in their studies of the cesium halides (7.27J, especially CsBr. In fact, they found vinually all the essential elements of what has come to be known as the Hartmann-Hahn method which we discuss later, though it seems evident that the full significance of the results of Bloembergen and Sorokin was not appreciated owing to the fact that they discussed the problem entirely in the Csl33 rotating frame instead of using the doubly JOtating frame which we take up in the next seclion. We give a brief account here of what they found, then return in the next section to discuss their results a bit further. using the concepts of Hannumn and Hahn [7.28] and of Lurie and S/ie/lter [7.29] concerning spin temperature in the doubly rotating frame. We shall discuss two important discoveries made by Bloembergen and Sorokin. 1be first has to do with spin-locking (Sect.6.6.1). We saw in Chap.6 that if the Ht is sufficiently strong (6.84) one must use Redfield's ideas of spin temperature in the rotating reference frame. Then, if one puts the magnetization along the HI in the rotating frame, it will persist for rimes as long as those for which the spin system may be considered to be thermally isolated from the surroundings. In their sample of CsBr, Bloembergen. and Sorokin found that the Cs t33 spin-Ianice relaxation time was lo-2Dmin at room temperature, ten thousand times longer than the Dr Tt 's. Thus they expected to be able to maintain (i.e. "lock") the Cs lJJ magnetization lined up along the HI)es for many minutes. Yet, when they auempted to observe the Cs l33 while sweeping through the line using an HI for which Redfield theory should apply. they found no signal even when they took only 15s to go through the line. They found that the explanation was that the short Br T t causes the Cs·Sr dipolar coupling to fluctuate in time, thereby providing a relaxation mechanism for Tie Eq. (6.9Ib). Since TIe specifies how long the spins can be locked. the Br T I thereby limits the Cs 133 spin-locking. A simple view of their discovery is that the Br T} contributes to the relaxation rates Tb and Tc of (6.87) and (6.9tb). The short relaxation time of the Br results from the modulation of the Br electric quadrupole interaction by means of lallice vibrations. Thus. we see that if one nuclear species has a dipole coupling to a second species with a fast spin-lattice relaxation time, Tie of the first species will be shortened, and one can detect the existence of the second species by comparing, for the first species, Tl with T( e' 275
Their second discovery was thaI if they utilized two oscillators. one tuned to the es l33 resonance, the other tuned near the 8r79 or BrS l resonance. they could utilize the rapid spin-lattice relaxation time of the bromine to polarize the es 133 . This effect they labeled the transverse Overhauser effect. In their experiment, they applied a strong rf field H de. tuned exactly to the Cs resonance. and a rather weak rf field HI )Br tuned to a frequency'; which differed from the Br resonant frequency VBr by an amount given by (7.80) As we explain rigorously in the next section, this condition can be understood by considering the effective fields Heff)Br and Heff)es in their respective rolating frames. In the bromine frame. HeffBr=' ) k
21r(v - VBr) 11l,
'H) IBr
+~
(7.81)
';;t
k2'IT(v - vBr)/rBr
The problem of matching the Zeeman splittings of two different species could be solvro if by magic one could apply one magnetic field to the I-spins, a second magnetic field to the S-spins. How can one do this to spins which are neighbors on the atomic scale? The magical solution was found by the Wizard of Resonance. Erwin Hahn and demonstrated by the Wizard and his Sorcerer's Apprentice Sven HartmiJnll [7.28]. Hahn recognized that alternating magnetic fields have a negligible effect on nuclei unless the frequency of altemntion is close to the precession frequency. Thus if he applied two allernating fields HdJ and Hds at frequencies WJ and ws, respectively, the frequencies being chosen to satisfy the respective resonance conditions, ~=n~
Since HdBr is small, Heff)Br
7.12 Hahn's Ingenious Concept
(7.82)
Likewise.
since the Cs is tuned exactly to resonance. The Bloembergen-Sorokin condition is a special case of the equation (7.83) which has come to be known as the Hartmann-Hahn condition. As we shall see, when this condition is satisfied, the two spin systems can come to a common spin temperature in the doubly rotaling frame. Equation (7.83) is a condition for rapid cross-relaxation in the doubly I'OIating frame. Since Bloembergen and Sorokin analyzed their experiment by means of the single rotating frame of the es 133 , they did not have spin systems which came to a common temperature in that frame of reference. They discussed the problem in tenos of an Ovemauser effect (indeed they called it the transverse Ovemauser effect). They showed that when (7.80) was satisfied. the es magnelizarion Me. and the Br magnetization M Br satisfied the relation
~=U~
(7~)
HIlJ would have negligible effect on the S-spins and Ht)s would have negligible effect on the I-spins. Each species could then be viewed in its own rotating reference frame. Figure 7.14 shows the two rotating reference frames. Note that the z-axis is the same in both frames. It is in fact the direction of Ho in the laboratory. In Fig. 7.14 the magnetization vectors MJ and M s are also shown. Note that in a general situation they do not lie along either the z-axis or the respective HI'S, though they might if special ways of preparing the syslem were used. For the case where an M is not parallel 10 its HI, it will precess about the HI in the rotating frame with precession frequency [}J or [}s given by ~=nHI)J
%=~HI)s
(7.8~
These equations suggesl immediately that we can make the precession frequencies match by adjusting the ratio of the H)'s to satisfy the relation HI), is ~~=W
n~
(7.84) In Sect. 7.16 we shall see how we can get their result in a simple manner by combining Redfield's results on the calculation of Meq, (6.9Ia), with the idea of spin temperature in the doubly rotating frame.
y,
,/' --",n/ , M,
\
...
__/
,
I
),"/
/l,p
flg. 7.14. The I'Olaling reference frames ahow lhe axes r, and rs along lhe respeelive rf fields Ill), and III )$' Note lhill the z-axis is in common
276 277
This condition is known as the Hahn condition. It corresponds to making the Zeeman splittings of the I-spins quantized along Hdr in the I-spin rotating frame equal to the Zeeman splitting of the S-spins quantized along H1)s in the S-spin rotating frame. Satisfying the Hahn condilion matches the Zeeman spJiuings, but in reference systems which are exotic to say the least Our real objective in producing the matching is to permit the two spin systems 10 couple. How can we use the exotic matching 10 couple the systems? We know that the spin systems are coupled via the dipolar interaction. For two spins of unlike species that portion which broadens the line is the A term of the dipolar coupling 1idA 1idA = 'Yr,}S h2IzSz(l - 3 cos 2 8rs ) T rs
(7.88)
Since the z-axis In the rotating frame is the same as in the laboratory, this coupling is unaffected by the transformation of spin variables to the rotating frame. (Note: Sometimes confusion arises as to what variables are transformed. In quantum mechanics one can formally introduce operators which transform only I, only 5, or only r, or any combination. Since r rs and 8r s are unaffected by rotations of spatial coordinates about the z-axis. (7.88) is unaffected whether or not r is transformed. As we show below, the usual approach transforms only I and 5 in order to remove the time dependence of the HI couplings to the respective spins.) Taking the classical view, if Ms is precessing around HI)S in the S·spin rotating frame, M~s will be oscillating sinusoidally at ils = 'YsHt)s. This produces, via 'J-idA (7.88), a time-dependent coupling to M:r. But M:r is the component of M r transverse to HI)r, hence the coupling produces the same effect as applying an alternating field along the z-direction of frequency ils. When the driving frequency as matches the I-spin resonance frequency il r, the S-spins cause the I-spins to absorb energy, and Mr to nutate away from the direction of HI) [. We have described two spin systems which are coupled. We know with coupled systems that when the natural frequencies coincide, resonant transfer of energy results and we need to worry about the back reaction. Thus for two coupled pendulums, one al rest initially, the other set in motion, after a while the first reaches a maximum amplitude, with the second one at rest. Then the energy exchange reverses, the pendulum which was driven now drives, the pendulum which drove now is driven. If we had only a pair of spins, a similar situation would occur. However, typically there are many coupled I-spins, many S-spins, perhaps coupled, perhaps not, but in different neighbor environments in any case. For such a case it becomes useful to assign a temperature 8/ to the I -spins in their rotating frame and a temperature 85 to the S-spins in their rotating frame. Under these circumstances, M r always poinls along HI)r and Ms always points along Ht)s, but 278
as energy is exchanged, the sizes of M r and Ms change. In equilibrium 8r = 85 , Causing 85 to increase produces a heating of 8r. We see, therefore, that there is a way of coupling two different spin systems so that they can exchange energy. We need now 10 do three things: l. 2. 3.
set the qualitative arguments on a firm quantum mechanical basis; describe the experimental steps involved in putting Hahn's idea to use; analyze the expected results.
7.13 The Quantum Description We star! by writing the Hamiltonian of the system in the laboratory frame. 'J-i = 'J-izr(t) + 'HZS(t) + (1{d)r / + ('Hd)SS + ('J-iJ)rs
(7.89)
where 1{zr(t) is the Zeeman energy of the I-spins. It includes both a static interaction with field Hok and a time·dependent interaction with the two alternating fields. It is most convenient to consider that rotating fields have been applied, instead of linearly polarized alternating fields. We have, then, 1{zr(t) = - "fr'il· [kHo + i(HI)r cos w:rt + j(Hlh sin W:1t
+ i(Hl)S cos w:st + j(Hlls sin w:st]
(7.90)
where W:1 and Wzs may be positive or negative, 10 represent either sense of rotation, and where
1= I:Ij
(7.90,)
j
is the total spin vector of the I-spins. The terms ('J-id)rs, etc., represent the magnetic dipolar coupling of the Ispins with the S-spins, and so forth. We now wish to transform to a rotating reference system. In so doing, we are following Redfield as discussed in Chapter 6. However, our problem is somewhat different from his since we have two rotating fields. We therefore wish to transfonn in such a way that we view the I-spins and S-spins in Iheir respective reference frames. The transformation is readily accomplished by introducing the unitary operator T defined as T = exp (iw~rIzt)exp (iw zs5zt)
I, = I:I,j ;'
where
(7.91)
S, = I:S,k
(7.92)
k
are the total z-components of angular momentum of the two spin species. We define a new wave function 1/;/ by the equation
(7.93,) 279
then substituting T-1ljJ' for
,po SchrOdillger's equation becomes
_!!. atjJ' ::; 'H'.// i
at
or
(7.93b)
where H' is a lransfonned Hamihonian. Explicit evaluation of 'H' using the techniques of Chapter 2 gives
'H!
= - 1'11/[(Ho + W:lhl)lz + (H ')1 IzJ
-1's1J[(Ho +w,shs)Sz + (H,)sSzl + ~Il + ~lS +~ss + lime-dependem terms we ignore . (7.94) The lenns 1t~11' elC., represent thai pan of the dipolar coupling 1tdl1 that commutes with the Zeeman inleraclion between the spins and the static laboratory field Ho. These terms are usually called the "secular part" of the dipolar interaclion. We write them OUI explicitly below. The lime-dependent terms are of two sons. One variety arises from the nonsecular parts of the dipolar coupling. They oscillate at frequencies Wl1. wzS. or w::/ ±wzs, The second sort arises from couplings of the [-spins to (HI)s and the S-spins to (Bd,- These oscillate al a frequency (wz/-wzS). Since wz/' WzS. and (WzI ±wzS) are all far from any of the energy level spacings in the rotating frame, they can be neglected. One must remember, however, that it is conceivable that should there be a quadrupolar interaction added, and should the two nuclei have similar ,'s (as do, for example, Cu63 and CuM), the frequency (Wzl - wzS) might, in fact, be close to a possible transition. If one chooses (7.95) where Hoo is a particular value of Ihe stalic field Ho around which we laler wish to make variations, then the two nuclei are exactly al resonance. Note that the two w's are negative if the 1"S are positive, representing the fact that nuclei of positive, rotate in the "negative" sense about Ho. Making use of (7.95), defining ho by Ho - Hoo '" ho, and neglecting the rime-dependent lenns of 1t', we have
1£' '" -1'rh[hoI: + (Ht),I:z] -1's1i[hoS: + (HdsS:z] +'H?t11
+'H~'S+1-l?tSS
and similarly for 'H~SS' To these may be added the pseudo-dipolar and pseudoexchange couplings when their size is large enough to be imporlanl. II is the tenn 1£3,5 which gives rise to the effects observed by Bloembergen and S.orokin [7.27] in their studies of CsBr. They found that the rapid bromine spin-lattice relaxation could communicate itself to the Cs nuclei through this lenn when the Cs nuclei were quantized along their own HI. We can view the various tenns of (7.96) as energy reservoirs of Zeeman or dipolar energy. Sioce the various tenns do not commute, they can exchange energy. Such processes may be tenned cross-relaxation in lhe double-rolating reference frame. The rate of cross-relaxation will depend on how the energy levels of the different tenns match, on the hcat capacities, and on the strength of coupling as measured by the failure of tenns to commute with one another. Thus, we note that the Zeeman tenn of the I-spins 1£Z! commutes with the Zeeman energy of the S-spins 'H.zs. However, as long as (Hlh /: 0, 1iZ! does not commU{e with either ~11 or ~,s' and we can transfer energy between 1izl and or Moreover, 1t~,S provides a coupling mechanism to transfer energy between 1tz! and 1tzs [provided (H tls /: OJ. All these remarks lead one, following RedM/d, to assume that if one waits long enough, the various pans of (7.96) will come 10 an equilibrium in which the system can be described by a common temperature 8. For some purposes, it may also be possible and convenient to assume that various parts may comc to common temperatures faster than the whole system achieves a single temperature. This is the viewpoint Lurie (7.29] adopts to calculate some cross-relaxation times. We therefore, make the assumption that when the system has achieved a common temperature it is described by a density matrix I! given as
1i311
exp(-l£'lk8) 'Ii'lk9)}
,= T, {e,p(
E = T, {,'Ii'} = _
(7.97)
(7.99)
where 1£' is given in (7.96). In lenns of I! we can calculate the average energy E and the average magnetization vector (M r) in Ihe high temperature approximation
(7.96)
It is convenient to define the Zeeman energies l£zl and 'H.zs by the equations
1i3,s'
C,[(H j )7 +
h5 + HCl + Cs[{Hj)~ + hij] 9
(M r ) = T, {e1r H } = Cr(~"'>r
(7.100a)
(7.100b)
where C, and Cs are the Curie constants given in tenns of the number of J. or S-spins per unit volume N J or Ns, and Boltzmann's constant k by (7.98)
C _ 1'yh2I(l + I)N , J 3k and where
280
etc.
(7.101)
HE is defined by the equation 28'
C,H1. • 0 • 9 = Tr {e( n dJ1 + 1tdlS + ftJ,SS) .0
.0
7.14 The Mixing Cycle and Its Equations }
(7.102)
.
Evaluating the trace gives 2
I
2
2
1'Y}NS S(S+l)
2
HL='j(Ll H)lI+(Ll H)/S+'j ,;N/I(l+I) (Ll H)ss
(7.103)
where (/1 2 H)OI!J is the contribution (in Gauss) ?f the. ,a-spins to the ,second moment of the a-spin resonance line. H L has the dimenSions of a magnetic field. Although we call it the "'ocal field", it should not be confused with the Lorentz local field. Actually, HL is introduced simply to enable us to factor Ct out of various equations. The fact thai the dipolar energy is (-Ct H[l8) makes it appear superficially thai we have taken into account only the I-spins in calculating the dipolar energy. However, thai such is nol lhe case is seen by examining (7.102) and (7.103) which exhibit explicitly the dipolar contribution of the S species 10 the expression for HL. C/He measures the tolal dipolar contribution 10 the spin specific heat. Note that although the term "local field" sounds vague, H L can in fact be calculated exactly and is to be considered throughout as a precisely predicted quantity. The only exception to the statement is found when pseudodipolar coupling becomes prominent as in higher atomic number elements. In that case one would need to know the magnitude of the pseudo-dipolar coupling tenns to make quantitative predictions. We can observe (MJ) from our oscilloscope photographs of the initial height of the free induction decay following tum-off of (HI)I' One funher expression is needed. It is the expression for the magnetization MI found after the demagnetization of I's. That is, suppose (HI)s = (H III = 0, and that M I = kMJO where (7.104) is the thennal equilibrium magnetization of the I -spins at the lanice temperature 9,_ With ho:> HL' we switch on (Hdl' and slowly reduce ho to zero. We then end up, according to (7.100) with (7.105) Note that if we were to change (H')I slowly, (Mf) would follow (HI)/, in accord with (7.105).
202
There are two general ways of doing double resonance. The first was proposed by Ha"rtmann and Halm [7.28]. The second, which is a variant of the first, was demonstrated by Lurie (7.29}. As will become apparent, the two different techniques have simple analogies in thermodynamics. Consider two bodies connected by a rod to provide thennal contact. One body, of small heat capacity, represents the low abundance S-spins; the other, of large heat capacity, represents the I-spins. Hartmann and Hahn's experiments are analogous to heating the large object by holding the small one at constant elevated temperature. The rate of heating depends on the thennal conductivity of the rod and the heat capacity of the large object (the I-spin system), but is independent of the heat capacity of the small object (the S-spin system) since we never let its temperature change. A theoretical prediction of the rate of heating of the large object would require knowledge of its heat capacity and of the thermal conductivity of the rod. In resonance language, that means we must calculate a cross-relaxation time. This cannot be done exactly. Lurie's experiment is analogous to breaking the thermal contact between the rod and the small object, heating the small object to a known temperature, disconnecting the heater, and reconnecting the rod. After a sufficiently long time, the entire system of large object, small object, and rod, comes to a common temperature. Since the S system has a relatively small heat capacity, the final temperature is not much different from the initiallemperature of the large object. However, we can repeat the cycle. In fact, if we do the thennal mixing N times, the heating of the I system is as great as it would be for a single mixing with an S system whose heat capacity is N times larger than it actually is. Since N may be made very large, a significant effect can be achieved even when the S-spins have a very small relative heat capacity. Calculation of the temperature rise requires knowledge only of the heat capacities of the pans. It is nOI even necessary for the heat capacity of the rod to be small since we can easily include its effect. Calculation of the heat capacities of the spin systems is simple and can be done exactly. We therefore have a simple. exact theory to compare with experiment. As Hahn and Hartmann's analysis shows, the effective thennal conductivity of the rod depends on the size of lhe two rotating fields. The Hahn condition provides the fastest mixing or largest thennal conductivity. The heat capacity of the spin system is detemlined in large measure by the strength of the HI '5. We can Iherefore vary the heat capacities experimentally although we must remember that when the HI ratio does not satisfy the Hahn condition, it may take a longer time for a unifonn temperature to be reached. The dipolar coupling between the two different species provides the thennal contact or "rod". As we have remarked, we can easily calculate its heat capacity. Likewise, there is a contribution to the heat capacity from the dipolar coupling of the I-spins among themselves and 283
•
!
the S-spins among themselves. All these effects can be rigorously and simply included. In the process, Lurie demonstrated that it is nOl necessary for the HI's to be large compared 10 the local fields and further demonstrated the coupling in cases where (H»/ has been turned to zero. We now tum to an analysis of the Lurie experiment. We shall assume throughout that spin-lattice relaxation can be neglected during the times of the experiment. Spin-lattice processes can be included readily, but one must ~ careful in so doing to include the sort of transverse Overhauser effects descnbed by Bloembergen and Sorokjn [7.27]. We begin by the demagnetization process. This brings (M/) down along the x/-axis, ils magnitude being given by (1.105). Let us call this magnetization (M/}j. During this process, since (Ht)S is zero, 11.zs commutes exactly w.ith the rest of the Hamiltonian. So, likewise, does 'YshSz. Therefore, (Ms) remains unaffected, and points along the static laboratory field Ho. The rest of 11./ is at a common temperature 8j which we can compute from (7.l00b) and (7.105): 0; = C/(H,)//(M/);
(1.106)
This temperature is, of course, very much lower than the lattice temperature 8/. We now tum on (HI)s suddenly. In such a rapid process, the state of the system does not change. The dipolar energy and the Zeeman energy of the I-spins are therefore unchanged. The S-spin Zeeman energy E s (1.101)
since (MS) and (Heff)S are perpendicular. The total energy of the system Ei is therefore
Eo = _ CJ!(H,)j + Hl1
8j
I
(1.108)
After a sufficiently long time, the S-spins Zeeman energy comes into thenna! equilibrium with the rest of the system at a common final temperature 8r· The final energy E r is then
_ CJ!(H,)j+H1J+Cs(H,)~ Er- 8r
(1.109)
But since the total system is isolated and its Hamiltonian independent of lime, its energy cannot change. Therefore, Ej=Er
(1.110)
This gives us that
8j 8r
_
--
where 284
C,(HI)l + Hl] 2 =-C,(H1)1+HlJ+Cs(Ht)s 1+£
(1.111)
(1.112)
£=
In the process the magnitude of (M,) drops from its initial value (M/)i to a final value (M/lr which, in view of Curie's law, is (M/)f/(M/); = 1/(1
+ e)
(1.113)
We now suddenly turn off (Ht)s. Once again the system immediately after the change has the same wave funclion as it did jusl before. The expectation value of 'Hz, and of the dipolar energies is thus unchanged, but that of 'Hzs is zero since (Heff)S = O. The tOlal energy E} is therefore
CJ!(H,)j + HlJ
(1.114)
Of
Immediately afler the tum-off of (H»s, (Ms) is nonzero. Therefore, we see that we do not have thennal equilibrium. If we wait for a sufficiently long time, the entire system will come to a common temperature 8n" In the process, (Ms) decays 10 zero. This is an irreversible decay. We shall in fact calculate the entropy increase below. When the system has reached the final temperature 8rr, the energy is Eff. Using (7.100a). we have
Ejr=-CJ!(H,)j+HlvOff
.
(1.115)
But Eff = E} since the spin system is isolated from the outside world and has a Hamiltonian independent of time. Therefore, using (7.114) and (7.115) (1.116)
8r =8ff .
Using Curie's law we see thai following the lurn-off of (Hils, (M/) does nOl change. For one complete on-off cycle, therefore, we can argue that (M,) is reduced by the factor 1/(1 + E). We can repeat the argument for another on-off cycle. The magnetization M/(N) after N cycles is thus given in tenns of its value M/(O) prior to the first cycle by
M/(N)/M/(O) = [1/(1 +e)IN When
£« I, as
(1.111)
in our experimenls, we can write this as
N!J(NVM(O) = e- N < where £ is given in (7.112). Double resonance is possible even when (H 1)/ «HL. Experimentally we accomplish this by perfonning adiabatic reduction of (HI)' after M, has been brought along (H I)' in the rotating frame. After (HI), ::::::0, /11, ::::::0 from Curie's law, however, we have retained the order in the I-spin system, the order now 285
being with respect to the local field [7.30]. (Hj)S is now cyc~ed o~ ~nd off N times. After the Nth cycle, (Hlh is adiabatically returned to Its onglOal val.ue. The resulling M[ is then observed by rapidly turning off (Ht)/ and observmg the free induction decay. TIle analysis for the case when (Ht h «. H L is essentially the same as given above; however, now, the tenn C[(H1)iI8 no longer appears in (7.108) and (7.109), and E: reduces to E:
=C5(HI)~/C/Hl
(7.118) 7
Lurje studied Li metal for which the 93 % abundant isotope Li gives the strong or I-spin resonance; the 7 % abundant isotope Li 6 is the weak or S-spin system. To observe the effect of the double resonance, he measured the amplitude of the Li 7 free induction decay afler N mixing cycles. Figure 7.15 displays Lurie's data demonstrating the effect of N on the destruction of the Li 7 signal. The solid line has no adjustable parameters apan from the N = 0 intercept.
60
.. '.... = '",," I f1/StC
50
o 1. . - t",= 1001/Stc
\ '1.~.
• • • • •
•
F1g.7.J6. In(Ah) versus (lldl II.l 1.5K (or N ..... 22, ,011. low 1,2, and IOms, (lId7 2.30G. For this value or (11.)7, (111)' =: S.tG salisfies the IIlI.hn condilion. The solid line is cakulated (rom (7.112) and
= =
=
o '.,.. - '"," 2mst>c .~
•
(7.117)
a
\. a
.'
\" 10023456
7 II U·1I 1 (gUlI$J)
')
10
Figure 7.16 shows how the spin temperature idea works even when the Hahn condition is nOl perfectly satisfied. Note that the solid line has no adjustable parameters. The deviations for larger H I )6 and shoner ton and tolr arise because there is not enough time for a common spin temperature to be established. In this example a very large destruction is produced. Alternatively, one can conclude that a much smaller fraction of Li 6 nuclei could produce an observable destruction. In this case one begins to worry about the long distance that energy must diffuse from the hot Li 6 spins to couple to the distant Li 7 's. This problem has been investigated experimentally by several authors [7.23.31.32].
7.15 Energy and Entropy
10 0
40
SO
120
160
200
Numbtf Qll'lIlst$ (N)
t·lg. 7.15. Lurit's experimentsl dll.ta showing In(M7) versus N at 1.5K. (l1d7 =:= 2..14~, (1I 1 )e = 504 G, nearly satisfying lhe lIahn tondillon. 4. =: tow =: 4 ms. The _lid IlIIe IS calculll.l«l usirl& (7.117) II.nd (7.118)
286
It is interesting to follow the changes in energy and entropy of the spins that take place during the double resonance. The essence of the experiment is the heating of the I-spins brought about by the contact with the hot S-spins. There is a net flow of energy into the system as a result of work done on the S.spins. The destruction of the I magnetization corresponds to an irreversible loss of order, that is, an entropy increase. The energy of the system is, of course, the expectation value of the Hamiltonian of (7.96). Basic theorems of quantum (and classical) mechanics tell us that the total energy remains constant as long as 1t does not explicitly depend on time. Rea.mngements of energy within the total system, even when 1£ is independent of time, we identify as a heat flow within
287
the system. Changes of the total energy due to varialion of an external parameter we call work on or by the spin system. We can follow the cycle by considering work and heatlfansfer in the rotating frame. Consider one complete cycle of (HI)S on and off. We stan by turning on (HI)S suddenly. Bearing in mind that the 5 Zeeman energy is
(HZS) = -(Ms)· (H,)s
(7.119)
and that during the sudden tum-on the dipolar energy does not have time to change, we see that it takes no work to tum on (HI)s since (Ms) is initially zero. The establishment of the 5 magnetization causes (1tzs) to go from zero to a negative value. That is, there is a heat flow from the 5·spin Zeeman reservoir to the rest of the spin system. That this is the direction of heat flow is reasonable since the initial zero (MS) in the presence of a nonzero (Ht)s can be viewed as saying that 1tzs has an infinite temperature. Associated with this heal flow between systems at different temperatures there musl be an entropy increase. Following establishment of (MS)' we tum off (Ht)S' Using (7.119) we can see that we must do positive work on 'H.zs in the process. [Note that during the tum-on or tum-off, which takes place suddenly, there is no time for heat flow, so that we can compute the work done solely from the changes in (1tZS) given by (7.119).) Following tum-off, (Ms) decays irreversibly to zero. Again there must be an entropy increase associated with the irreversibility. We are now ready to repeat the cycle. Note that we have done a net amount of work on the spin system, and thai there has been an irreversible loss in order. Had we turned on (Ht)S in the next cycle before (Ms) had been able to decay, the S·spins would have done positive work back on us. In fact, had (Ms) not decayed at all, we would have gotten back as much work as we put in when we turned off (H ,)s. We would not then have done any net work in a cycle. Moreover, apart from the effects at the original tum-on, there would have been no irreversible loss in magnetization of either spin system. We must allow a sufficiently long time for the irreversible features to occur. The entropy 0" of the system can be calculated starting from the basic equation u =
E+k81nZ 8
(7.120)
where Z is the partition function. Making use of the high-tempemture approximation, we can evaluate (7.120) to get 0" ':l!'
k[ln(Trl)-
2k~82 Tr~~2}]
(7.121)
Evaluating the Tr {'}i2}, as in (7.100a) gives (with Ho = 0)
C/[(Ht)~ + HlI + Cs(H I 20'
)1
(7.123)
(Note that (7.105), describing the adiabatic demagnelization, follows from use of (7.IOOb), Curie's law, together with the requirement that the entropy given by (7.123) remain constant.] Since we have worked out the temperature at each part of the cycle, we can use (7.123), together with the approximation that E: is small, to find that in one complete cycle starting at temperature 8 there is a total increase in entropy (1r - O"j of ur-O"i=
CS(Ht)~ 82
(7.124)
Half of the increase occurs following the tum-on, the other half following the tum-off at (HI)s. We note that the larger (H tls, the larger the change of enlTOpy per cycle. Since the existence of M, is a sign of order, we see that a large (H t)s leads 10 a large destroction of M" as was, in facl, already expressed by (l.118).
7.16 The Effects of Spin-Lattice Relaxation In Sect. 7.11, we saw that Bloembergen and Sorokin observed IWO striking ef· feclS in their pioneering double resonance experiment. As we have mentioned, they gave a full explanation of their resullS. We tum now to an ahemative description of their experiment to show how one can think of their experiments in teons of two concepts: (I) spin temperature in the doubly rotating frame and (2) Redfield theory of the equilibrium spin temperature reached in the rotating frame when spin-Iauice relaxation is included. Thus, we show that their experiment can be viewed as a combination of Redfield's experiments on saturation with the Hartmann-Hahn and Lurie·Slichter experiments. As we explained in Sect. 7.11, Bloembergen and Sorokin observed a species, [, (Cs 133 in their case) which had a long T\. The (H I )/ they used was such that one should consider the I·spins to be described by a spin temperature in their rotating frame. They observed that for the [-spins, TIl? was deteonined by the T1 of the 5-spins, leading to a circumstance that (T'e)/ «(TI )/· They observed further that if they were on resonance for the I-spins (7.125)
"IJHo=w/ but off resonance for the S-spins such that
where Tr 1 means the total number of states, and equals Tr 1= (21 + I)N'(25 + I) N s
(1 = k[N/ In (2J + I) + Ns In (25 + 1)]
hSHo -
wsl
= ''YJ(H\)/ with
(7.126')
(7.122)
"Yj(H,)~«(-ysHo -wS)2
(7. 126b)
they produced an I magnetization, M" given by 288
289
•
I
MI = ±M/(O"Ho)1s , (7.127) 1/ where M/(ot. Ho) is the I magnetization when the I-spins are in thermal equilibrium in a static field Ho with a lauice al temperature (J/. To understand how (Tt)s can determine T 1g for a case in which (Hils = 0, we follow the Bloembergen-Sorokin analysis. We consider a case in which (HI)j::> where HL is some sort of local field. Then the I-spins. if exactly al resonance, are quantized along the I-spin x-axis, the direction of (H I )/. and the S-spins are quantized along the z-axis, the direclion of Bo_ The term 'Hd/S • given by (3.55), involves products such as I;k5zp. which have matrix elements which are diagonal in Szp. bUI are off-diagonal in J;J;p- However, owing to the spin-Ianice relaxation time of the S-spins, the diagonal matrix elemenlS of Szp are functions of time with a correlation lime of (Tl)S' Defining
we can set
Ta ! = (T1 )! for the S-spins we will have. in analogy to (7.132),
aMzS Mos - M"s --/)t
HE
Ckp=
"'fnSh2
3
rkp
2
(I-3cos 6kp )
(7.133)
-
with
TIIS
(7.134)
and
(7.135)
oMzs Mzs ~=-TbS
(7.136)
where TbS ...... (TI)S. We will have three dipolar relaxation equations similar to (6.87):
~(7i )_ fJt dfJ -
(7.128)
( 7idll) T,11
(7.137a)
BlQembergefi and Sorokjfl show that
_'_=.2..l: C 2 (TI,,)1
1i 2 k,p
5(5+1)
kp
3
(T,)s
I +wrl(TI)~
(7.137b) (7.129)
The reader should compare this expression for the relaxalion of (Iz ) along (HI)z with the expression for TI of (5.372) which applies to the relaxation of (I,,) along Ho. Physically, the fact that the z-component of the S-spins is fluctuating produces a time-dependent magnetic field on the I-spins along the z-direction. A field of this polarization is ttansverse to the I-spin quantization direction, and can therefore relax the I-spin magnetization along its "static" field in the rotating reference frame. The factor TI/(I +wr,Tls ) gives the spectral density at the frequency WII which can produce flips of the I-spins quanlized along (HI)!. If (HI)s is turned on to a level sufficiently strong that the $-spins obey the Redfield spin-Iemperature condition in their rotating frame (6.84), then the magnitude of the S-spin magnetization, Nls, and of the I-spin magnetization, M" will be given by an argument such as is given in 5ecI.6.9. The calculation we described above for (Tle)l can be viewed as a calculation of Tb' defined by a modified (6.86):
z / = - M:z:d oM-
at
'I
w'lh ,
( 7idlS)
where we have neglected the tenns like (~) of (6.87). If we assume that (Tt)s <: (Tt ), the lattice will flip individual S-spins much more rapidly than individual I-spins, so we can neglect any rates dependent on I-spin flips. Therefore (7.138) and both will be much shorter than Tell - Likewise
TIl& <:Tb/ -
0 ) E = - (ho)M: s - (H,)/M:r./ + (lldfJ
+ (7i?'/s) + (~SS)
The z-component of M, will not be relaxed directly by the time variation of 1-$ dipolar coupling since that coupling commutes with I:. Thus in the equation
290
(Til)!
.
(7.140)
Then
aE
at =
at
(7.139)
In analogy with (6.91) we will then find for the Bloembergen-Sorokin case
(7.130) (7.131)
aM:J - = MOl - M:J
(7. 137e)
TeiS
(7.132)
I oM:s - (IO)S---al -
aMzl
(Hd,---at
a"~ a"~ a 0 + at ("dll) + a,!"d/s) + at (7idSS )
(7.141)
Keeping only the fast tenns which relax at rates comparable to (Tt)s, we get 291
•
I
BE = -(ho)S (Mos - M.s) _
at
(1i~15)
T as
_
(1i~SS)
TclS
We now assume the system has a common spin temperature
M
M
(7.142)
TeSS
B so
that
- Cs(ho)s 8
(7.143a)
- ClCHill
(7. 143b)
zS -
:d -
()
o
(1i dIS ) = -
CsHJs e
(7.144.)
(7.144b)
where Hys and H~s are defined by (7.144). To find the equilibrium spin temperature, we set dE/dt = 0 and substitute (7.143) and (7.144) into (7.142), gelting
()
Mos(ho)s/(ho)~ + Hl ls + Hl ss ) Tas
TaB
Tcls
(7.145)
TeSS
If one is off resonance (lto)s by a sufficiently large amount compared to the local fields, one can keep only the [enn involving (ho)S in (7.145), getting
Cs = Mos (J
iJ
1(1'0)51 MOl (ho)S
(7.146)
(lto)s
BUI, since
Ms = Cs(ho)s
8 MS=MoS
(7.147.) (7.147b)
Recognizing that
CsHo Mos=--
8,
(7.148)
(7.150)
The ratio l(ho)sl/(ho)S is either +1 or -1 depending on whether ws lies below or above resonance respectively. The two possible signs reflect the fact that 8 is positive in the fonner case [Ms parallel to (Hcff)s] and negative in the laHer [Ms antiparallal to (Heff)s]. The size of MJ is larger than its thermal equilibrium value in the ratio ishJ' This result is just the result of the Overhauser effect, hence the name "transverse Overhauser effect". In considering this result, we see that it follows very directly from the realization that if the S-spins are far from resonance
Ms =Mos
(1t~ss) = - Cs:~s
Cs =
Ml = "
(7.151)
and that the I-spins and S-spins are at a common spin temperature in the doubly rotating frame. We could achieve the same result by an appropriate adiabatic demagnetization experiment. We tum on (HI)J and we tum on an (HI)S which satisfies the Hanmann-Hahn condition at a frequency Ws far from resonance. We then sweep ws onto resonance. The two spin systems will then mix. However, since they have comparable heat capacity, the common spin temperature they reach will be holter than the temperature the S-spins would reach if the spins did not mix. So we tum off (H.)s, go off resonance. magnetize the S-spins again, and repeat cooling the I-spins several times more. In this way, we can bring the I-spins asymptotically 10 the value given by (7.150). However. we can only hold this result for a time (TI/?)! given by (7.129). The Bloembergen-Sorokin method gives one a steady-state MJ along (HI)J' Note that if one sits with the S-spins exactly at resonance [(ho)s = 0], (7.145) and (7.146a) show that the equilibrium values of Ms and MJ would vanish even if one satisfied the Hartmann-Hahn condition. Thus, to use the fast relaxation of the S-spins to generate a strong I-spin magnetization, the S-spins must be far from resonance (above or below). Equation (7.145) can be viewed as giving the spin temperature of the Sspins. If the Hartmann-Hahn condition is satisfied, the I-spins will rapidly come to this spin temperature. However. if one is far from the Hartmann-Hahn condition, the I-spins will not reach this temperature, as is shown from the experiment of Lurie et al. (Fig.7.16).
we get, combining (7.146) and (7.147), M = C/(Hl)/ = C/(Hlh CsHo = (Hdl MOl r 8 Cs(ho)s 8, (ho)s
(7.149)
7.17 The Pines-Gibby-Waugh Method of Cross Polarization
But from (7.145)
'"a (HilI = 'Ysl(1to)sl so thai we gel the Bloembergen-Sorokin result that 292
The methods of Hartmann and Hahn or Lurie et al. involve detection of a rare species by observing its effect on an abundant species. Pines, Gibby, and Waugh [7.33] pointed out that under some circumstances it is preferable to perform 293
the experiment in a manner to observe the rare species. Specifically, suppose C 13 is the rare species (its natural abundance is 1.1 %), and HI is the abundant species. as when one is working with solid hydrocarbons. Each nonequivaicni C 13 site will have its own chemical shift. One would like to record each C l3 chemical shift. We have seen (Sect. 5.8) thaI the Fourier transfonn method is then a very efficient way of collecting data. since a single C l3 7r/2 pulse will excite all the C 13 lines simultaneously. If onc observes the HI resonance, one can only measure the individual C13 resonance lines using a point-by-point search. However, if one brings the HI magnetization along (Hdll. the hydrogen HI, one can rapidly polarize the C 13 nuclei by turning on their H I> (HI )e. to the Hartmann~Hahn value. Within a cross-relaxation time, Ihe C I3 will be polarized. Since there are many more hydrogen atoms than C 13 nuclei, in this process the HI spin temperature will hardly change. Therefore
Me
-- GdHlk
o
(7.152)
But
8 = Gu(H1h,
and
(7.153.)
Mil
1'dHde = 111(Hd"
Me = M u Ge Gn
so thai
(I'll) 1'e
(7.153b) (7.154)
Suppose. now, Ihal initially MI-I has ils {hennal equilibrium value in Ho
GIIHo 0, The corresponding MC<8/) is given by MH = Mil (0/) = - -
(7.155)
Ho MdO,) ~ Ce9;
(7.156)
Ulilizing these equations in (7.154), we find
Me ~ MdO,) 711 7e
(7.157)
This is the Bloembergen-Sorokin enhancement, characteristic of an Overhauser effecl. Having polarized Ihe C13·s, one turns off Iheir H" and records the free induction decay. Meanwhile, one leaves on the proton HI for two reasons: (1) to maintain the prolon magnetization [spin-locked by (Hl)lI] and (2) to provide so-called H 1_C13 spin decoupling. This last topic we take up later in the chapter, where we see that the presence of the strong H I acting on the protons effectively wipes out the HI splillings of the C I3 resonance. By this means the C I3 resonance consists of a family of lines each of which goes with a given C13 chemical shift, but without splittings arising from Ihe C I3_H I spin-spin coupling. 294
We now lum on (HI)c again, repolarizing the C13· S along their HI. We again tum off (Hde and record the C l3 free induction decay. Each repetition of this cycle will heal the HI nuclei in accord wilh (7.112) and (7.117), so that the C13 signal gets progressively smaller until it becomes necessary to tum off (HI)}! to allow Ihe proton spin-lattice relaxation time to replenish the HI magnetization. We turn now to a few remarks about the nature of the rare spin spectra. If we recall that in general the spin-temperature approach requires that we be working with solids. we should distinguish single crystals from powders. In a single crystal a given type of C I3 chemical site might have several different bond orientations with respect to Ho. For example, in a benzene molecule (C6H6), Ho might lie along one CH bond, in which case il would also lie along one other in the molecule, but would make a 60" angle with the other four bonds. Thus we would have two C13 lines, with one twice the intensity of the other. If the sample is a powder, the C 13 spectrum of this bond would be a smear. By magic-angle spinning (discussed in Sect. 8.9) one could narrow this pauem to a single narrow line at the average chemical shifl position of C I3 'S in Ihal molecular position.
7.18 Spin-Coherence Double Resonance -Introduction The Overhauser-Pound and the cross-relaxation methods of double resonance are both dependent on the existence of relaxation processes of sufficient vigor. In the case of Overhauser-Pound schemes it is necessary that appropriate spin-lattice mechanisms are Strong. In the case of cross-relaxation double resonance. crossrelaxation times must be sufficiently rapid (relative to spin-lattice relaxation) for the method to work. We now tum to the third family of double-resonance schemes. For these schemes to work, Ihere must be a coupling between the two spin systems which manifests ilself in splillings of the spectral lines which would exist in the absence of the coupling. We observe the double resonance by producing some fonn of modulation of this coupling. Crudely speaking, we observe the effect on splittings of the I-spin resonance produced by exciting the S-spin resonance. If 8w is a typical such spin-spin splitting. the essential condition for viability of Ihis third family is that the effects of Ihe splitting not be obscured by relaxation processes. If one thinks of 8w as some sort of beat frequency, Ihe period of the bealS is 1/8w. The requirement is then physically that the coherence of this beat nOI be interrupted by relaxation. If T stands for the relaxation time (spin-spin, or spin-Ianice) which could interrupt the coherence, observation of the coherence requires that T:p 1/6w. The requirement of long relaxation times shows that spin-coherence double resonance tends to work under circumsr.1nces where the other methods do not.
295
7.19 A Model System - An Elementary Experiment: The S·Flip-Only Echo To begin the discussion of spin-coherence double resonance we are going to imagine ourselves to be working in resonance before the invention of all the varied double resonance methods which exist today. We suppose that we have observed a HI resonance which is split into a doublet by coupling to another nucleus (for example a C 13 ). Since the C 13 affects the HI spectrum, we ask ourselves is there not some way we can use the HI resonance to detect when the C l3 is brought to resonance? We first describe the model system, a pair of coupled spin ~ nuclei I and S with a Hamiltonian similar to that of (7.29) 1{ ""
-"nhHo1, -1shHoS, + 1-£12(t)
Initially we shall ignore relaxation effects, considering 1-£12 to be independent of time. For simplicity we take it to be (7.158)
This is the fonn of the indirect spin-spin coupling in liquids. If I and S represent different nuclear species, the tenns I~5~ and I1I 5 11 are nonsecular. Then we approximate (7.159)
This is also the fonn which the dipolar coupling between two nonequivalent spins takes if we keep only the secular pan, see (3.55). For indirect spin coupling in liquids, it is conventional to substitute J for A, and to speak of the J coupling. Since we wish to include dipolar coupling between unlike nuclei, we utilize A as the coupling constant. If I and S are the same species but have different chemical shifts, we can represent the chemical shift difference by defining (7.160)
where 1 is characteristic of the species. If then 1(71 - 7s)Hol:> A
(7.161)
we can approximate 1{12 by (7.159), but if the chemical shift differences are comparable to A or less than A, we must keep the complete expression AI· 5 for the coupling in a liquid. For a dipolar coupling in a solid, we would need to include both the A and B terms of (3.7). At this point we make a diversion to discuss notation. In the literature it is customary to refer to the case where (7.161) is satisfied as an "AX" case, whereas the case where the full expression (7.159) is needed as an "AB" case (do not confuse the A of AX with the A of AI· 51). Each letter refers to a given chemical shift or resonance frequency. 296
1t "" -1IliHo[, -1sliHoS, + AI,S,
(7.162)
The energy levels are
.
'Ys "" 1(1 - O's)
Letters next to each other in the alphabet have chemical shifts comparable to their spin-spin couplings. Letters not adjacent have big differences in resonance frequency compared to their spin-spin coupling, either because the chemical shifts·are quite different, or because the nuclei are different species (HI versus C I3 ). An A2 system has two nuclei with the same chemical shift. An AB2 or an AX2 system has three nuclei. In the fonner case the chemical shift differences between the identical pair are not large compared to the A-B spin-spin coupling, in the laller case (AX2) they are large. A notation AMX means three nuclei with big chemical shift differences. Note that if two nuclei are different species, they would be called an AX system. We shall treat an AX system, hence take
E""
-hwO/ml - hwosms + AmimS
(7.163)
where
WOI
-= 11HO
(7.164.) (7. 164b)
If we apply alternating fields at frequencies WI or ws (near to the resonance frequencies of the I-spins or the S-spins, respectively), the resonance condition i, (7.165.)
WI =wO/-amS ws "" wos - ami a
-=
where
(7. 165b) (7.166)
Alll
!'
Since ms "" ± the I resonance consists of two lines, spaced apan in angular frequency by a, as does the S resonance. Thus, through the couplings the presence of the S manifests itself in the I resonance. We now ask the question: Is there some way we call use this manifestatioll to ulllble us to employ the I resonallce as a means of detecting when we are producing a resonallce with the S-spins? In fact, in 1954. Virginia Roydell [7.34] was interested in just this question in order to measure precisely the ratio of the 1"S of HI to CU. We describe her experiment and that of Bloom and Shoolery [7.35] based on the theory of Bloch [7.36] in Soct. 7.20.
One simple concept immediately comes to mind. Suppose the S-spins were polarized so that they were all in a given state ms, say ms "" +~. Then the I spectrum would occur at WI "" wOf - aI2. We now stan searching for the S resonance by applying Jr pulses to them of angular frequency ws. We slowly sweep ws. After each S-spin pulse, we inspect the I resonance by inspecting 297
I-spin ~bslH"ption
(,)
Fig. 7.17. I-spin absorption spedrum for S-spins 100% l>olarizcd with (a) "'5 = and (b) "'s vcrsua I-spin scar<:h frcqucncy 1<1/
= -t
+t
before
~fter
(,)
Ibl
n
1.111 .1/2
w'I·~/2
",
I (J
n
~bsorplion
(bl 1.1 01
,,
,
n ",
wtI· a/2
Idl
I_spin
ms: 1/2
1.1 11 - a/2
-1/2
1.1'1' ~/2
n
",
wOI-~12
1.1 01. a/2
",
, l
I
(e)
the I absorption speclrum. It will be unchang~ un/~ss ws was tuned to th~ S resonance. In that case, we change ms from +~ to -'2" and the I resonan~e shifts from WI = WOI - oil to WI = WOI + 0./2. The situati?n is illustrated in Fig. 7. 17 . Simple and beautiful as this scheme may seem, It ~Ias cannot be ~sed exactly as described without first gelting the S-spins polanzed, a very difficult .~k indeed. In fact, at nonnal applied field Ho and temperatures 8, Ihe condillon "Ys hHo/ k8 <. I holds, so that to a good approximation one has an equal number of spins ms = as ms = We can reanalyze the experiment very Simply by saYing tha: half of the Ihe ot~er half have mS = -'f.' Let u,~ call I-spins have neighbors with ms = these two groups (1) and (2) respectively. Figure 7.18 shows the before and "after" spectra for the two groups produced by the 11" pulse on the S-spins. Group (I) will have a resonance at WI = WOI - 0/2 before the S-spins are flipped, and a resonance at WI = WOI + all after. Group (2) will have a "before" re~nance S·spinS are at WI -- w01 + a/2 ' which will be shifted to WI = WOI - all. after the .• flipped. We observe of course the spectrum of all the [-SPinS, which IS the sum of the spectra of the two groups (1) and (2). As we see from Fig.7.18e,f, the total spectrum is the same before and after. It looks as though our scheme has failed. . Fortunately, though the simplest version of our concept does not work, It can be modified very simply in any of several methods to be made to work. The essential point is that anyone [-spin belongs to either group {l~ or group (2). Since eirher group (I) or group (2) is affected by flipping the S-spms, both. group (I) and group (2) can tell the difference between whether or not the S-~PInS are flipped. We must do an experiment which utilizes the memory that SpinS have
+l
-l·
.
+l,
298
.
If!
n
n
1.1 01 - ~12
1.1 01.
al2
n
",
wOI-~12
n ",
1.1 01 •
~/2
Fig. 7.18. The I,spin absorption before (lcrt column) and after (right column) applying a 'II" pulse to the S-spina, for the group (I) l-spil1$ (a, b), the group (2) I-spina (c,d), and the total of alii-spina (~,f). Note that though the s.-::trum of ~ilhu group (I) or group (2) alone i$ changed by Ripping the S-spins, the loIal intensity pdtem (~.f) is unslTec:l.ed
of their prior history. An experiment which displays a spin's memory is easy to concoct. At t = 0 we apply a pulse to the I·spins with a sufficiently strong (HI)! to flip the spins in both the I absorption lines. We view the I-spins in the reference frame rotating at wO/. In this frame, we will have two magnetization vectors of equal length, one precessing at angular frequency af2, the other at -0./2. If we assume that immediately after the 7((2 pulse the I magnetization lies along the y·axis [Ihe situtation if (Hj}1 were along the x-axis], the magnetization will obey
trn.
(Mi('»
= C;i (cos (at/2) + i sin (a'l2)]
+
~oi [cos (ain) _
i sin (ai(2)]
= COi(eiat/2 +e- iat / 2)
2
Co = (MI(O,»
where
(7. 167a) (7. 167b)
(7.168)
gives the thennal equilibrium magnetization of the I-spins for the lattice tern· perature 8,. 299
Fig.7.20a-<:. p31 NMR signal versus time for diethyl phosphite I(C2IhOhPHO). (a) The precession of the p31 nuclei due to lheir coupling to the IP nuclei; (b) and (c) the precession of the p31 nuclei with the HI nuclei inverted al the indicated times. The dashed lines in (b) and (c) show whal the precession would have been had lhere been no II! inversion. (Data taken by Charles Pennington wilh assislance from Je3n-Philipl>C Anscrmel, Dale Durand, and David Zax.)
Then, at a time t1f we apply a 7r pulse to the S-spins. The group (1) spins will instantly change their precession frequency from WOf - a/2 to WOf + a/2 and the group (2) spins from WOf + a!2 to WOf - a/2. Thus, both groups will reverse their precession directions in the rotating frame. (Readers who are not content with the present informal treatment can look ahead at Sect. 7.24 for a fonnal justification of the relationships we use here.) We can utilize the complex notation to express this rotation reversal. If a vector is represented by a complex number A with initial components At: and Ayobeying Az =
Ao cos 4>
A y = Ao sin
4>
(7.169)
where Ao is the magnitude of the vector, then A = A z + iA y = Aoexp(i4». If this vector is rotated at angular velocity w, it becomes at a time t later A(t) = Ae iwt = Aoei4>e iwt
o
(7.170)
Therefore, we can apply this relationship to the two components of (Mt(t» given in (7.167b) to get the magnetization at a time t - t.,.. after the S-spin 7r pulse as
(Mt(t» = ~oi (e iat "/2 e -ia(t-t,, )/2 + e-ial"/Zei«(!-t,, )/2)
(7.171)
Thus (Mly(l» =
Co cos (at.f2 - a(t - t.)f2)
(7.172)
At time t - t 1r = t 1r (t = 2tll")' (Mfy(t» will once again be Co. In other words, we have produced an echo of the I-spins by flipping the S-spins. The Sspins so-to-speak cause the I-spin "field inhomogeneity" (a two-valued function). By flipping the S-spins we reverse this field inhomogeneity. The theoretical time development of (Mfy(t» is shown in Fig.7.19. An experimental demonstration of this phenomenon by Pennington et al. from the author's laboratory for the p31 resonance of liquid (CzHsOhPHO is shown in Fig. 7.20.
A useful way of viewing this experiment is to plot the phases 4>1 (t) and 4>z(t) of the two groups of I-spins where we define their phases at t = 0 as being given by (7.169). From (7.171), they are 4>,(1)=
(Tr/2)
4>,(t) =
o
It. S_spins inverted
300
C'f2J+ atf2
+ atll"!2 -
a(t -
t n )!2
C<(2) - «tf2 (7r!2) - at 1r / 2 + aCt - t n )/2
fOI
t
fOI
t> t1f
fOI
t
< t 1f
fOI
t
> t>r
(7.173)
These are shown in Fig.7.21. If there were more than one paicof spins I-S, each pair with its characteristic chemical shift, the situation would be more complicated, but the same principles would apply. Let us label each I-S pair by k. (k = I to N if there are N distinct pairs.) We shall assume the I-S coupling is only within a given pair. Then we have N Hamiltonians of the fonn Fig. 7.19. The y·componenl of the I-spin lllflgnetizalion versus time, t. Al t t", the S-spins are in\'erled by " 1f pulsc
=
1ik = -liWOfkIzk - flWOSkSzk
+ AkIzkSzk
(7.174)
so that, defining ak = Akl!l, 301
PhilSe iogln 4'lltl iod 4' l ll)
(7.176)
,
{Mj(O}""" = C;i l)exp [- i(!7 lk - (lk!2)t]
+ exp [- i(!7 lk + ak!2)t] Fig. 7.11. The phllSe
(7.177)
This general expression as well as the effect of a 'Ir pulse applied to the S-spins at time t" can be represented by plotting ,pu:(t) and q,2k(t) given by
for t
0,,(1) = (./2) - ({Ilk - a,/2)'
(7.178) for
t > t"
A similar expression holds for 92(/.;) except that the sign of Qk is reversed. In Fig. 7.22 we plot the ,p's for a case of two I-S groups (N = 2). It is convenient to give a name to the experiment we have been analyzing. We shall call it the
PlliU ¥lgle
S-jlip-on/yecho.
7.20 Spin Decoupling
Fig. 7.22. Thc phll$C IIngles ~1 k(t) and ~H(I) versus time, t, for a ca.sc with lWO 1-5 pllit'll (k 1 and 2). At 1 I., the S·spins lire nipped by 11'. Al t 'It:r, ~u(t) ,pH(I) for bosh spin groups, indicating thllt bolh components of I magnelization for II given pair sre collinear. Nole, however, lhal since lhe two pairs have differenl chcmical shifls, lhe I magneli."aliOllS of lhe different pain are out of phase wilh each other
=
=
=
=
= Coi L(exp [ _ i(WOlk '- ak!2)t]
2
,
+ exp { - i(WOlk + ak/2)t]
(7.175)
in the laboratory. Relative to some frame TOtating at WI we have magnetization (Mj(t)"",s' Defining 302
The ideas we have just developed lead in a natural way to the concept of spin decoupling, one of the earliest goals of double resonance experiments. We will follow a pedagogical rather than a historical order. The goal of decoupling is to simplify spectra. A typical NMR spectrum consists of many lines arising from the combined effect of chemical shifts and spin-spin couplings. Decoupling is a process which effectively eliminates the spin-spin couplings. It is very important both for high resolution spectra of liquids and spectra of solids, for example to eliminate the effect of proton spins on C I3 spectra. We shall discuss removing the spin-spin coupling between different nuclear species such as HI and CU. The basic idea on which all decoupling schemes work can be understood physically as follows. The existence of two lines in the I-spin spectrum corresponds to the fact that the S-spins have two orientations, up and down. If we can cause the S-spins to flip back and forth between the up and down orientations sufficiently rapidly, we should achieve an effect much like the motional narrowing of resonance lines. That is, an I-spin will precess at a time-averaged frequency rather than at one or the other of two discrete frequencies. To study decoupling, we are going to pick a scheme which is easy to analyze. We will look at the effect on the I-spins of applying a sequence of 7r pulses to the S·spins. We shall see that as the time between S-spin 7r pulses gets shoner and shorter, the I-spin spectrum goes from two distinct lines separated by Q in angular frequency, to a single line at the average frequency. There are two general approaches we could lake to showing this result. One would be to try to calculate what the I-spin absorption spectrum Xl/(w) looks 303
like as we flip the S-spins al a progressively more rapid rate. Or, we could look at Ihe Fourier transform of the absorption line, which we do by looking at the time development of the I-spin magnetization following a rr{l pulse applied to the I-spins. In the absence of S-spin pulses, the I-spin magnetization will consist of two components, one oscillating at WOI + 0/2, the other at WOI - 0/2. So, if we choose Wf = WOf. in the Wf reference frame the components will precess at angular frequencies +a{l and -0/2. If the system were perfectly decoupled, a would become effectively zero, so both components are at rest in the WOf reference frame. and the transverse I-spin magnetization will be a constant in time. (We are not including relaxation effects in our Hamiltonian. If we did, the I-spin transverse magnetization would decay rother than remaining a conSlllnt in time.) We shall follow the Fourier transfonn approach for whi<.:h, in facl, we have already set the srage in Secl 7.19. There we analyzed what happens if we apply a single 'lr pulse to the S·spins al a time t = t .... Let us then look back at Fig. 7.19 and at (7.171), which show that at time t = 2t,.. the magnetization (Mt(t)} has returned to its value at time t = 0+, immediately after the 1r{l pulse. lllOugh Ihe magnelization al the time of the 11" pulse is smaller by the factor cos (at ...{l), the 1( pulse has produced an echo al t = 2t.... If we now think of ourselves as having a fresh start at t = 2t ... , we realize we can repeat the echo by applying a second 'K pulse to the S-spins. If we apply il at a lime t~, hence t~ - 2t... after the first echo, we witt again fonn an echo. It will occur at a time 2(t~ - 2t ... ) after the first echo, or at a time til given by til = t~
+ (t~ - 2t... )
(aJ
F1g.. 7.23 (a) The magoetiution (M,,(t» versus I for the CAlle of a ,.. pulse applied to lhe S-spins at I producing a 21 .. , followed byanrefocusing al t other,," pulse at t I~, with a sc<:ond refocusing at time I". (b) The effect on (M 'J'(t)} of applying a pair of ... pulses to the S-spins at times I.. and I = 31 .. , with nsulting echoes at t 2t.. and
=
>t.
= /'" =
=
(b)
k
'.
II
21.
lal
(7.179)
see Fig.7.23a. If we chose t~ to be given by t~ = 3t...
then
(7.180) (7.181)
so Ihat the time delay of the second echo after the first is identical to the time delay of the first echo after the initial signal. We can repeat this process again and again. The situation is shown in Fig. 7.23. The phase angles ;,(t) and ¢2(t) of Fig.7.21 will then appear as shown in Fig.7.24a. If we choose a shoner time for t,.., as shown in Fig.7.24b, the maximum excursions of ¢l (t) and ¢2(t) away from tr{l will be less, so the magnetization versus time will appear like the solid curves in Fig. 7.25 mther than the dashed curves. Clearly, (Mfy(t)} will be a periodic function as long as there is no relaxation, with period, T, equal 10 2t,... We can think of (Mfy(Y)} as consisting of a constant with a superimposed periodic ripple. Figure 7.25 shows that the shoner t ... , the smaller the ripple, and the more (Mfy(t» approaches a constant in the reference frame rotating at wOf, hence behaves as though a were zero. Thus, we have "decoupled" the S-spins from the I-spins. 304
Ibl
• ./2
i'E~-::"':;:'--:7<E:~-::"':;:'--:7i;<E:--,
".
Jo'ig.7.24. (a) The phl\.Se angles .pI (t) and .p,(I) of the two components of I-spin magnetization. The,," pulsoes: applied to the S-spins at I t .. and 31 .. cause.pl and .p, to become ~ua.l at I = 21 .. and 'It,.., correspo~ding to the maxima in the curves of (M,.(I)} shown m Flg.7.23b. (b) The effect on
=
=
305
Fig. 7.25. The effect on (Mfr(l» of halving the Lime I ... The solid-dashed curve corresponds to I .. of Fig. 7.24a, the solid-only curve corresponds Lo Fig. 7.24b. Note that ('""ffr(t» (!Ill be Vlewed as a steady value with a superposed ripple, that the ripple is periodic with period 21 .. , and that the amplitude of the ripple diminishes as I.. is shortened
/
We can eltpress the eltistence of the ripple mathematically if we wish. by recognizing from Fig. 7.25 that since (M/y(t)) is a periodic function with period 2t r it can be eltpressed as a Fourier series
=
L
(MJy(t» =
All cos (21fftt(J')
(7.182)
with
We can summarize for the case thai we flip the S-spins sufficiently frequently: following a 7r/l pulse applied to the I·spins. they precess as though the coupling to the S-spins were turned off. This method of decoopling was in fact proposed by Freeman et al. in 1979 [7.37]. They call it "spin-flip" decoupling. In their paper, they discuss its advantages over OIher methods, as well as discussing interesting variants. Another method of producing such a zero time-average z-component is to apply a constant HI at the S·spin resonance which will cause the S-spin to precess about the HI' As a result. the z-component of 5 oscillates sinusoidally at angular frequency 'Ys(HI)s, This method is historically the first method employed by Royden (7.34] and by Bloom and Shoo/ery (7.35]. To analyze it. we talee the Hamiltonian to be 1(. "" -
- ')'sMH ds(Sz cos wst - Sy sin WSt)
n=O
(7.183)
T = 2t.". The coefficient of the constant tenn, Ao. then turns out to be A = (M (9
o
J
I
» sin (at 7l"/2)
(7.184.)
at"./2
which shows that as atrn shrinks. Ao approaches (M/(UI»' The higher coefficients are An
:=
2{A1j(U,) sin (at r /2) (_ t)n+I atrn
(at r {l.)2
An cos (21rnt/2t,..) = An cos (at/2)
(7.186)
We have included the Zeeman interactions of the two spins, their spin-spin coupling, and the interaction of a rotating magnetic field (Ht)s with the S-spins. We have nof included (he interaction of (H ds with the I-spins since we assume that Ws is close to the S-spin resonance, but far from the I-spin resonance. We now view the problem in a double reference frame which is the lab frame for the I-spins and one rotating at ws for the S-spins. That is. if the original Hamiltonian ?i and wave function 1/1 obey (7.187)
(7.I84b)
(:lI"o)2 - (at r n)2
which become much smaller than the constant tenn as at r - f O. Note that if atr{l. is very large., so that there are many oscillations in the I-spin rotating frame between S-spin flips. the coefficients An are very small for small n. However, for :lI"O = at r /2. the coefficients peak. (For larger values of n they decrease again.) Substituting into (7.182) this gives a time dependence for such values of 0 of
where ?i is given by (7.186). we define a transfonned wave function transfonned Hamiltonian ?if by Vl=e~iwstS.¢
•
l/J'
and
(7.188)
which. when substituted into (7.187), gives
_~ al/J' i
at
= ?i'.I,f
with
(7.189)
¥'
(7.185)
This value corresponds to the I-spins precessing at +a/2 and -a/2 in the rotating frame. the situation without dccoupling. The transition from the split lines to the strongly dccoupled lines will therefore look much like Fig. F.3 in Appendilt F. Strongly decoupled spins require that there be no peak in the An's at frequencies .:::: ± an. hence that in (7.185) there be no tenn for n ~ I which has a denominator close 10 zero. This condition requires (atrfl) < 11". If we look a Fourier transfonn of (7.182), we would get a constant tenn in the reference frame rotating at WOJ, and side bands at frequencies ±(n/Ur ), where n is an inleger. The amplitude of the side bands diminishes relative to the constant tenn as one shortens t r . 300
"tJItHoI: - 'YshHoS: + AI;S;
H' =
- ~1"Ho[. - ~s"{[Ho - (wshs)]S. +(H\)sS.)
+ AI;S:
(7.190)
Defining (ho)s == Ho - (wshs)
and
(7.191.) (7.191b)
we get
?i = -')'JhHOI; - ')'!l(Heff}S' S + AI:S;
(7.192) '<)7
This Hamiltonian can in fact be solved exactly, but we will leave the exact solution as a homework problem. Instead, to solve this equation, we assume thai we have applied a strong decoupling field so that (7.193)
(m/mslIIz:lm~ms') = (m/IIz:lm'/)om m' s' S' so that the energy changes aE are
A.E(ms') = "UltHo - Ams' cos
(J
(ho)s
and treat the tenn AlzS z as a perturbation. It is then convenient to define axes (x', y', Zl) such that Zl lies along (Heff)S' and the y- and y-axes are coincident (Fig. 7.26).
,
•
= "'1lhHo - Ams',=~~~~
";(ho)~ + (Hl)~
(7.199)
±!.
Corresponding to the two values of ms' = there are thus two I·spin resonance tines whose spacing is a function of how far ws is tuned away from the resonance frequency wos. At exact S-spin resonance, (ho)s = O. the I-spin resonance consists of a single line. Figure 7.'2:7 shows how aE(ms') - 'nhHo goes with (ho)s/(HI)S, demonstrating the collapse of the splitting as one tunes the S-spins to resonance.
, FIg.1.26. The effective field acting on the S-spil"llll in their rotating frame as a result of (lids nd of being some..·h...t off-resonartee for the S-.pins ((ho)s ct OJ
(7.198)
A
FIg. 7.21. Spin decoupli"8' the ener&>' splitting, i1E - '"f,U/ o , seal in the I-spin res0nance as a fundion of (ho)s the amount, the S-llpin ill off resonance, showing that the two transitions collllp6e to a single transition as the S-spins approach their resonance condi· tion (ho)s = 0
An
This gives Sz = Sz' cos 8 - Sz:! sin 8 with
(7.194.)
r~(;;;ho:;)g,S=""
(7. 194b)
cos 8 =
J(HI)~ + (ho)~
The unperturbed Hamiltonian
Ko =
-A/2
-A
?to' is then (7.195)
--YJhHOlz - "YS(Heff)SSz'
with eigenstates lm/ms,) which are eigenfunctions of I z and Sz" The perturbation AlzS z then becomes AlzSz = Alz(Sz' cos (J - SZI sin 8)
.
(7.196)
But Ihe tenn in SZI has zero diagonal element in the Im/mS') representation, giving for the energy eigenvalues
E'mJm S' = -""f11IHoml - "Ysh(Heff)sms' + Am/mS' cos 8 .
(7.197)
A weak rf field tuned close to the I-spin resonance can then induce transitions with a selection rule given by 308
The exact solution of (7.193) introduces another feature, the fact that the direction of the effective magnetic field acting on the S-spins depends on mI. the orientation of the I-spins. This same problem arises in electron spin resonance and is treated in Chap. 11, (11.83-88). There it is shown that as a result the frequency of transition for .1ml = ± I is modified, and in addition transitions nonnally forbidden in which ms' changes become possible. Decoupling by applying a steady altemaling field was. as remarked previously, demonstrated by Royden [7.34J and by B/oom and Shoo/cry [7.35] based on ideas of Bloch {7.36J presented in an invited talk at an American Physical Society meeting. In Fig. 7.28 we show data from Bloom and Shoolcry, demonstrating experimentally the collapse of the splitting as the S-spin oscillator approaches resonance. 309
.'lg. 7.28. Dcmonstntion by Bloom and Slwolery of the effect on the F19 resonance of varying the p31 resonance frequency for an (11 dp of 800 Hz. L}v is the amount (in Hz) that the p31 is off-resonance. As L}v approaches zero, the splitting goes to zero. "lh~ ill the top line denotes (fft}p
Noise Modulalion
Jll1111ul1J Jl zoo Ht
"" "'
400 Ht
"" "' "" ",
~~~l J~ Jt .At JJIIJJ II11 700 Ht 800 Ht 900 Ht
No'M;;;;; lar
No double
,esona~
Fig. 7.Z9. Comparison of the dceoupling with random noise and with a coherent rr frequency. The FIB spedrum of CHFCl z is observe
As is evident from Figs.7.27 and 28, the collapse of the splittings is only perfect when the S-spins are exactly on resonance, and the size of the splitting grows linearly with the S-spin offset for small offset. As a result, it takes a large (H,)s to deal with a situation in which there are several S-spin chemical shifts present in the spectrum. Anderson and Nelson [7.38] and Freeman and Anderson (7.39] employed frequency mcxlulation of (H\)s to improve the decoupling over a broader range of (1I o)s. Ernst proposed a method of broad-band decoupling (7.40] in which he used a noise signal to modulate an rf carrier, for example, by switching the rf phase between two values (0 and 11") at random time intervals. Figure 7.29, reproduced from his paper, shows a comparison of noise modulation with an unmcxlulated (Hds as a function of the S-spin offset for the case of CHFCI 2 in which he observed the F 19 spectrum while decoupling the protons. Clearly noise modulation produces a dramatic improvement in the ability to decouple when not perfecl1y at resonance. This result follows from the fact that with a sufficiently broad noise spectrum, the residual splitting varies quadratically with offset instead of linearly as in Fig.7.27. Some other decoupling schemes for broad-band decoupling are a coherent phase alternation method described by Grutzner and Santini [7.41], in which a 50% duty cycle square wave is used to phase modulate (H,)s. Basus et al. [7.42] discuss the advantages of adding a swept frequency for ws to the coherent phase shifting. The approach of Freeman, Kempsell, and levitt can be further improved by the use of socalled composite pulses. Invented by Leviu and Freeman [7.43], they consist of a closely spaced group of pulses to achieve a desired rotation (such as a 1f pulse) 310
which is insensitive to some parameter such as the frequency offset of (HI)s. For example, instead of an X(7l") pulse, they suggest and analyze an X(1r!2) followed immediately by a Y(1I") followed immediately by another X(1I"!2). The theory of composite pulses has been reviewed recently by Levitt (7.44] and by Shako and Keeler [7.45]. Freeman has given an excellent review of the status of broad-band decoupling (7.46].
7.21 Spin Echo Double Resonance We turn now to a slight variant of the basic double resonance experiment described in Sect. 7.20. The technique was first demonstrated by Kaplan and Hahn [7.47] and by EmshwiUer et aJ. [7.48], and is called spin echo double resonance (SEDOR). This technique is important because the observation of a SEDOR proves that the two nuclei involved are physically near each other. Consider an ordinary spin echo with the echo fonned at 2T. During the first time interval T, the spins dephase, but during the second iOlerval T they rephase. Hahn pointed out that if a nucleus of spin I has a neighboring spin belonging 10 a different species, an S-spin, the neighbor produces a local field which may aid or oppose the applied field, broadening the resonance much like a magnet inhomogeneity or the existence of several chemical shifts. Of course, magnet inhomogeneities do not affect the echo amplitude since the dephasing effect of the inhomogeneity during the first interval T is exactly undone by a rephasing during the second interval T. He noted, however, that if one added a second oscillator to flip the S-spin with a 11" pulse at the time the I-spin is given its 11" pulse, the sign of the field the S-spin prcxluced at the I-spin would be opposite in the two time intervals T. Thus, if a neighbor dephased the I-spin during the first interval, it 311
_ 50~~-~~--'~~~-~~
Fret/lIt'IIC)' of Clf 1800 pl,lse mell$lfred ,e/atil'e 10 Clfi) mllill line (k1I~)
• r-'.~12~.,----_----,~~.,---__~4~.,---~ 20
~ 1.'3
'."§ ~
j
• • • • •
......•
•
40
60
•
•
'ooL
.,:-__---,4~.~ ,, ,
..
.... ....
:
'•
.;
.'
•
o Spin Echo
.~ 35
~ 30
,
~
"";;25
~20
§ PoSitiol' of 0,·)
z
,
.~
•
..
•
o
•
0 0
0
~ ••
5
• " ' ,•
.L--'-£-,--L-c--J~~',---'-.-,+c__--'-~J. 1.07
1.08
U)9
mallllllle
---'---
• SEDOR
, :
: \
~ 10
1.10
1.11
1.12
1.13
1.14
1.15
1-10'*'0 (kG / MH%)
---.J
Fig. 7.30. Boyce's observation of SEDOR in • Cu powder containing O.~ at. % Co near the Cuu resonance fT«luence of pure Cu at 9901 Gauss and 1.5 K, showing the satellite due to the 1 .... _ 1 transition of lhe first neighbor. The parameters UBed are T = 250"" 1I is :::::5Ga:" for ~"If pulse of the Cu transition.. IOO echoes 'llere averaged for each point and the dots are larger th.n the S(:a-ller and dnft. Note that t.he ?"uch more abundant Cut3 nuclei a long distance from the Co, the SOotalled Cu malll hne, do nol show up in the Co SEDOR
t .... - t
would continue to dephase during the second interval, producing a smaller echo at 2T. An example of a resonance delected in this manner is shown in Fig. 7.30. Boyce [7.49-51) studied a dilute alloy of Co in Cu. He wished to observe the nuclear resonance of Cu nuclei which are near neighbors of the Co nuclei. For such a dilute alloy, the neighbor resonances are weak in amplitude. frequently hidden under the tail of the resonance of Cu nuclei which are distant from impurities, the so-called "Cu main lines". Boyce was able to reveal the "hidden" resonance of the first neighbor by doing a spin echo double resonance in which he observed the effect on the Co spin-echo amplitude of a 1r pulse applied over a sequence of frequencies close to the Cu main line. Since distant nuclei do not produce much field at the Co, the "main line" has negligible effect on the Co echo height. But the first neighbor has a big effect. In this instance the use of double resonance can be seen as a way of selecting pairs of nuclei which are close in space. Makowka et al. [7.52] utilized pt195_C13 SEDOR to detect the surface layer of Pt nuclei for small metal panicles of Pt on whose surface they had adsorbed a monolayer of CO enriched to 90 % in C 13 . The metal particles. which were tens of angstroms in diameter, were supported on A1203, a typical supported catalyst. Makowka et al. observed Pt'95 spin echoes. Figure 7.31 shows their data. The straighl Pt 195 spin echo gives lhe line shape of the Pt nuclei in the small metal particles. This line is over 3kGauss wide! To observe the Pt195 resonance of 312
, 4.
>-
g 15
•
• •
80
• 45 '2
Fig. 7.31. Measurement ofMaki:Jwla et at (7.521 of the spin·echo (0) and SEDOR (.) line shapes of Pt lU for a sample of small particles of Pt metal supported 011 J\12~ whose surface is coal.ed with C 13 0 molecules. The metal particles have diameters of a few tens of A. The Pt lU spin echo gives the totlll line shape of all the Pt ltS in the particle. The SEDOR data, involving pt 1U _C 13 0 double resonance, and lin add-subtract method, give the NMR line shape of the PllU nuclei in clo5e proximity to the CO molecules, i.e. the surfll.Ce layer of Pt atoms
the surface Pt atoms only, they employed an add-subtract technique in which on alternate spin echo cycles they applied a C 13 'If pulse coincident wilh the Ptt95 'If pulse. For those PI nuclei far from the C t3 , the echo was unaffected by the Ct3 'If pulse, whereas for the surface atoms, the Ptt95 echo was diminished by the C I3 pulse. Thus, subtracting the Ptl95 echoes when the C I3 pulse was applied from those when il was nOl, the signal from the Pt l95 nuclei not bonded to C t3 nuclei vanishes. To analyze the SEDOR signal, we nOte that the precession angle. 8, of the I-spins off resonance by (1I o)J during an interval T is
8 = (")'(ho)1 - ams)T
(7.200)
where (ho)1 represents lhe extent to which the particular I-spin is off resonance due either to magnetic field inhomogeneities or chemical shifts. Suppose, then, that at t = 0 we apply a 1r(2 pulse with (H')I along the I-spin x-axis in their rotating frame. This puts the corresponding magnetization M«ho)J) along the +y-axis (Fig. 7.32a). At time T- just before the I and S 1r pulses, 8 has the value 9, (F;g.7.32b):
8, = 80 =f a-r(2
where
(7.201)
80"" ,/(ho)iT
The I-spin 1r pulse reflects M«ho)J) about the x-axis, producing the situation shown in Fig.7.32c. During the next time interval, T, 8 advances an angle 82 (Fig.7.32d) given by 313
(a) t: O'
(b) t :T-
,
7i(I,5) = -'I"[(l>o),I, + (HI),I.) ~ l's rl[(h o )sS: + (HI)SS~J
,
M
M
e,
(7.205b)
" 7i(I) + 7i(S)
where by 1i(f) [1i(S)] we mean a Hamiltonian which is a function only of the spin components of the I-spins [S-spins]. If this Hamiltonian acts from a time t1 to a time t2, we have
, (e) t
(7.205,)
(d) to 2T
,
,
'f(t,) = ex p ( -*7i(I, 5)(t, - 'I»)'f(tl)
(7.206)
Since 1i(f) and 1i(S) commute, we can write this as t/J(t2) = ex p (
,, , c ,,I a
l
,
,
:~,
" : e" ,
M
=ex p (
FIg. 7.31a-d. The I-spin magnetization for spin echo double resonance (SEDOR) seen in the I-spin rotating frame, M is the magnetization component off-resonance by (h o), for a particular valu!) of At ! 0, (lIJ), is applied along the ~-axis, producing a 11:/2 rotation of M onto the y-axis (a). During the time r, M precesses through a net angle 91 given by (7.201) (b), At t r, 11" pulses are applied along ~I and ~s to both the I-silins and the S-spins, so that at ! = T+, M is at the position shown in (e). During the next interval, T, M precesses through the angle 92 given by (7.202), producing the situation of (d)
=
=
(7.202) where the upper signs of (7.201) and (7.202) go logether, as do the lower signs. Thus M«ho)/) makes an angle .c18 (Fig. 7.32d) given by .c18 = 81 - 82 = ~aT
.
XJ(.fl) ... T
(7.203)
til)ex p ( -*1i(l)(t2 - til)tJ>(t1)
. (7.207)
...
[XJ(')'XS(')]'" (' -
T)
(lIo)s, to allow for chemical shifts, and use the notations of (7.176)
fh "'" l'J(h o ll -
WI
ils = Is(ho), - WS
(7.208)
Then,
(I+('» =Tc{I+,(t)} (7.204)
It is useful for future reference to derive this result using the density matrix formalism. But first we demonstrate a useful theorem. While the (HI )'s are on, we customarily neglect the spin-spin coupling, giving in the doubly rotating frame a Hamiltonian 314
-~1i(S)(t2 -
1i(S)(t2 - til) t/J(tl)
We shall assume the I-spins are off resonance by (hO)1 and the S-spins by
Since Ll8 is independent of 80 or (Ito) I, it is independent of magnetic field inhomogeneity or of the existence of multiple chemical shifts. Since ms is equally likely to be either +~ or -~, two resultant magnetization vectors, equal in length to MOI/2 form, making angles d8 of aT or -aT with the negative y-axis. Thus
(My (2T» = -MOl cos aT
til)exp (
The significance of (7.207) is that in analyzing the effect of the rf pulses, at a given time, we can treat the effects of (HI)I independently of those of (HI)S' Thus, we can treat first the I-spins, then the S-spins, or vice versa. We do not have to treat both spins at once. Of course, the conclusion of this theorem is only valid during the time we can neglect the spin-spin coupling. We now employ this theorem to analyze SEDOR using the density matrix. We consider a sequence in which the I-spin and S-spin pulses are defined in the doubly rotating frame. At t = 0- the spins are in thennal equilibrium. At t "" 0 we apply a 7r/2 pulse with (H I )/ along the I-spin x-axis. We denote it as an X/(7r/2) pulse. We wait a time T, then apply simultaneously X I (7r) and XS(7r) pulses. We then follow the density matrix e(t) and the magnetization at later time, t, such that t > T, looking for an echo at t - T = T or t = 2T. We denote this sequence with the notation
'
M
"'s.
-*
-~1i(l)(t2 -
.I
(7.209)
To find U(t), we must first know its value at U(O-), corresponding to thennal equilibrium. We thus take
- =..!..(l
0 <><)Z+
1iWOIIz+1iwosSz)
kT
(7.210)
315
I h II
with Z the partition function. In calculating (1+(0) we need keep only the teon
,,0 ) _
'!WOt
We thus have
(I
+
(t» =
Tr{I+e- iat ,S,2T(_1 )eiat ,S,2T} ( 1+(2T) = hwOI ZkT y
(7.211)
= ZkT1,
+ . ZkT Tr{1 exp(l(flt1~ + flsSz - a1~Sz)(t - T»Xt(7I")Xs(7I")
fIWO!
x exp (i(fl!1~ + flsS~ - a1~S~)T)X t(1r(2)1z X J 1(ir/2)
The ql,lantity in the curly bracket is just what we would have if (ho)t and (ho)s vanished, and the system developed solely under the influence of the spin-spin coupling, a1zS z . Thus. chemical shifts and field inhomogeneities are refocused, but not spin-spin couplings. l'sing the mjmS representation to evaluate the trace, and recognizing that
x exp(-i(flt1~ + flsS~ - al~S~)T)Xsl(1r)XJI(1r) exp(-i(fltl~
x
+ flsS z - a1zSz)(t - T»)}
(7.218)
(mjmsII+lm~m~) (7.212)
Utilizing that
=
(~II+I-1)6msm~6ml,I/26m~,_1/2
(7.219)
we get 1
Xt(1r!2)I z X/ (1r!2) = 1y
Tr {I+e -illt.S.2T Iyeiat.S.2T}
(7.213)
and insening
= 1
X s I (11' )X J (If)X t(lf )Xs(7I")
(7.214)
before Xt(7I"/2) and its inverse after X J 1(1r/2) we then transfoon
L: (~II+I ~ !)eiamST (- !msllyl!ms)eiamST
m,
I I
But
I y = ~(r - 1+)
X t (1r)Xs(1r)exp (i(fl[ 1~ + flsS z - alz S z )T)Xs (lf)X/ J (71") = exp (-i(fl t 1z + flsSz)T)exp (-ialzSzT)
(7.220)
and
(7.221)
l
(7.215)
(~Ir-I-~) = ( - W-I~) = I so that
(7.222)
and
Xt(1r)XS(lf)1yXSI(1r)XJJ(1r) = Xt(lf)1y X
r (lf) = -Iy l
.
Substituting, we then get
+
hWOj
In the high temperature approximation, Z '" 4 so that
+.
(I (t) = ZkT Tr{l exp(l(flt1z + DsS~ - alzSz)(t - T))
(1+(27) = - hWkOt i cos aT . 4·T Using for thennal equilibrium
x exp(-i(flt1z + DsSz)T)exp(-ia1~SzT)(-Iy) x exp(ia1z S z T)exp(i(fl t 1z + flSSz)T) x exp(-i(flt1z +flsS~ -
a1zS~)(t
- T»}
.
(7.217a)
This expression can be simplified immediately since the teons exp (WsS~(t -T)) and exp( ~WSSzT) on the left of (-1y ) commute with everything between them and the teons exp(iflsSzT) and exp(-WsSzU - T» to the right. Hence, they can be combined, giving a factor of unilY. We then get (I+(t» =
~~~ Tr {1+ exp (i(Dt I z X
(7.223)
(7.216)
MOt = xoHo
')'2h 2 I(I
XO =
'Yfl(Iz}
+ I)
we get liwOj
exp (-Wt1zT)exp (-ialzSzT)( -1y)exp (ialzSzT) .
(7.225)
3kT
(I,) = 4kT " 10
a1zSz)(t - T))
x exp(+Wt1zT)exp(-i(flj1z - a1zSz)(t - T))}
=
so that
(1+(2T)} = -i1o cos a7
(7.2I7b)
At t - 7 = 7 or t = 27, the factors involving flt combine to unity, leaving
(7.224)
.
(7.226) (7.227)
The -i factor indicates that the net spin lies along the negative y-axis. We can now utilize the fonnalism, in particular going back to (7.212), (7.213), and (7.215), to compare three pulse sequences, the S-flip-only echo of Sect. 7.19, the conventional spin echo and spin echo double resonance:
316 317
I
S-flip-only echo (Sect. 7.19): Sequence: X I (7r(l) ... 1' ... XS(7r) ... (t - 1'). Conventional I -spin echo:
is applied at a time T after the I-spin 7r(l pulse (T S 1'). This sequence gives. in the notation of (7.201), Dl = 00 '=F aT(l ± a(1" - T)(l
82 = 00 ± a1"(l
Sequence: X I (1f(l) .. _1' ... X/(1f) ... (t - 1').
and
so that
iJ.0 = 8 1 - 82 = T aT
(7.228a) (7.228b)
and
(7.229)
([+(2T)) = -iI, cos aT .
(7.230)
SEDOR,
For these we need to insert for (7.214) respectively Basic S-flip only:
X
S'(lI')Xs(l'I")
Conventional I echo:
X/'(l'I")X I (1'I")
SEDOR,
X
s
'(ll')X/'(ll')XI(ll')XS(7r)
so that at time 21' we will have had for the transformation of the teon exp [i(ilII:+ nss: - aI:S:)1'} Basic S-flip only:
exp [i(ill I: - ilsS: + aI:S:)1']
Conventional I echo:
exp (i(-ilrI: + ilsS: + aI:S:)1')
SEDOR,
exp [i(-nII: - ilsS: - aI:S:)1') .
As in the transfoonation from (7.217a) to (7.2l7b). we can combine the teons involving exponentials of ilsS:(t - r) or flSS:1' on the left of (-I,,) with their inverses on the right of (-I,,). What counts are the tenns involving I: or 1:5:. Thus we see that: The basic S-flip-only 1'1" pulse changes the sign of the a teon but leaves the sign of the il/ alone. Therefore, it refocuses the spin-spin coupling at t = 2r but leaves the I-spin chemical shift alone. The conventional echo 11" pulse changes the sign of both spin-spin and chemical shift terms. Therefore it refocuses both spin-spin and chemical shifts at t = 21", eliminating both. The SEDOR 1l' pulses change the sign of the chemical shift tenn only. Therefore they refocus the chemical shifts, at t = 21', effectively eliminating them. but leaving the spin-spin coupling untouched. A useful variant on the SEDOR pulse sequence was introduced by Wang et a1. [7.53.54). The time of the I-spin 7f pulse is held fixed. The S-spin 7f pulse
This sequence has two advantages. The first is that one can work at very short times. T. while keeping the I-spin echo delayed a convenient time. This feature eliminates I-spin amplifier recovery problems. The second advantage is that the decay of the I-spin echo with pulse spacing (by T2 processes) is held constant as T is varied, and thus eliminated from the result.
7.22 Two-Dimensional IT Speelra - The Basic Coneepl We now arrive at two-dimensional Fourier transfonn magnetic resonance. one of the most exciting and powerful developments since the original discovery of magnetic resonance. In this section we introduce the concepts using a very simple example, the basic S-flip-only experiment discussed in Sect. 7.19. We shall find that it already contains all the essential ingredients of two-dimensional magnetic resonance. The idea of two-dimensional Fourier transfoonation is due to JeenU. He first presented it in lectures at the Ampere International Summer School. Basko Potze. Yugoslavia, in 1971 {7.55). An excellent review by FreemlJn and Morris [7.56] describes the explosive development of the field, and states and documents that "Ernst was the first to appreciate the great potential of the method". Indeed, both Freeman and Ernst with their collaborators provided many of the fundamental applicalions. Aue et al. gave a treatment of basic significance and broad generality [7.57] which also gives a detailed treatment of concrete applications. It is a fundamental reference for workers in this field. Recent reviews include those by 8ax [7.58} and Turner [7.59}. We shall describe in later sections some of the panicular fonns of double resonance, but OUf goal in this section is to introduce the concepts using S-f1iponly double resonance (Sect. 7.19). We saw in Sect. 7.19, (7.177), that for a spin I off resonance in the frame rotating at WI by ill =wOI-wl
the magnetization was 318
319
(Mt(t» '" i~O(e-i(01+0/2)t+e-i(01-0/2)t)
(7.231)
of strength
(M/(t» '"
(where the -fl , means the sense of rotation is negative) for t < t... and
(Mt(t» '"
+ i;Obe-iOatle-iOut2(eibI2/2 +e-ibt2/2) (7.232)
i~Oe-iO/(21.)
and we would get
lal
(2t~, 2t:. elc.) 2t'lr
<,
<
<,
(t - 2t~, etc.)
<,
,11 ""'-'_,.J-
_
".
2t.
t.
<
v-
lei /
/
"
v-
<, <,
,,
/
(7.233)
< Idl
I I I I I I
/ /
In terms of t, and t2, we have that for times after t"
(Mt(t» '"
Ibl
/
<,
r...
tl '" 2t'lf
<
/'
This, as we remarked at the end of the previous section, means that at t '" 2t.... the magnetization is just what it would be if the spin-spin coupling were zero. so that only the chemical shift fl , '" WO/-WI remains. [We here assume a unifonn static field, so that (ho), arises because of chemical shifts.] The situation is illustrated in Fig. 7.33 where we show (Fig. 7.33a) that for limes greater than 2tll"' the phases 4>1 and 4>2 act as though from 0 < t < 2t'lr there were only a chemical shift, but for t > 2t1l" both chemical shifts and spin-spin splittings exist. In Fig. 7.33b we have redrawn the situation to represent the situation of a '" 0 for 0 < t < 2t'lf' H we reset t'lf to a new value. longer than t'lf' the new situation is shown in Figs.7.33c and d. We can plot phases for a sequence of values t,.., t~. t~. etc., as in Fig. 7.33e. These plots can be summarized by defining two times. tt and t2 • by
t-
(7.237)
=2t .... we then have
(Mt(t» '"
t2 '"
i~Oa e-iOJ&t'e-in/&t2(eiot2/2 + e-iaI2/2)
i~Oe-iOlt
x (e+iat./2e-io(t-t. )/2 + e-iat./2e+ia(t-t. )/2) for t > t'lr' Exactly at t
b. we could label the chemical shifts as fl/a, n,b. and then
,.•
I I
2t~
21~
i~o e-inltle-inJt2(eiat2/2 + e- iot J/2)
" !(tj,")
(7.234)
This expression is derived for t) ~ 0, t2 ~ O. If at this point we treat this theoretical expression fonnally as existing for -00 < tt < +00 and -00 < t2 < +00. we can take its Fourier transform with respect to the two time variables tl and t2, to obtain I +~ g(W\,W2) == (211-)2 e-iwltle-w;1!2 !(tl.t2)dt, dt 2 (7.235)
lei
<
J
-~
Utilizing Ihe relation 6(x) :: (1/211") L~:: exp (ixt)dt. we get
ieo g(wj,"")' 26(w, + il,)[6(..., + ill - a!2) + 6(..., + ill + a!2)J. (7.236) These dala can be represented as points in a two-dimensional plot (WtoW2) (Fig.7.34a). If there were a second chemical shift and a second spin-spin spliuing.
320
2t~
2t~
2t~·
.'~g. 7.33a~ P~1lSC:S of the two I-spin magnelization components Yersus time for the SnIp-only ex~rllnent. The two phases ~I(t) and 412(1), showing the time t. aL which the ... pulse lS applied lo the S-spins, and the formation of the echo at t '" 2t which refocuses the spin-.pin splilling. H the spin-spin coupling were zero, the ~'s wo:k1 both follow t~e dashed curve of slope 0 1 rq>resenling a chemica.! shift only. (b) The ~hll.vior of. eqUlVlllent, for! > 21 .. , to that of (a). (e), (d) Replot.1ll olea) and (b) f('$pectively for .. value t~ >~". (e) A family of phue trajectorir$ colTesponding to a family of experiments with different times I. (t~, I~, de.)
321
IbJ
la)
0. ,, I I I
-IO~-bI21
--~-------""1-
-t QI,·bI21
-----------t-II
I
iI
---+------t--
iI
I
-!n I -aI21 ---~--
_ 0,
-ttl1u12 1
I
,, ,
I I
-Q l ~ -Q h -Q~~ Fig. 7.34. (a) The lwo-dimensional plol of frequencies WI and W2, 'lhowing poinL5 in "'I corresponding to the chemical shin WI = -(af ±a/2). (b) A plot similar lo (a) corresponding to two I-spins, each coupled lo a single S-spin with corresponding chemiCl\1 shift!! aid and nf6 l\nd spin-spin couplings a and b
iCO a 9(Wl.W2) = -2-5(wl + n!a){5(W2 + !1-la - 0/2)+ 5(W2 + n la + 0/2)] iCOb
+ T5(w, + n ,b)(5(W2 + n,b - bIl) +6(w, + [h, + WZ)J
17.238)
which would give {he plot of Fig.7.34b. This plot allows us {o go along the WI-axis {O detennine the values of the I -spin chemical shifts, then at each chemical shift position (o observe in the W2 direction splitting of those panicular I -spins by the S-spins to which the)' are coupled. That is, there is a correlation of the WI frequencies with the W2 frequencies. The physical origin of this correlation can be traced back to the phase diagram Fig. 7.35, which shows that the single phase line over the time interval 0 10 2t7: splits into two lines for limes after 2t7:' For t > 2t7:' those
lines extrapolate back to an intersection at t = 2t ... The two lines are diverging from their intersection since the chemical shift frequency fl/ is split into two frequencies. n/ ± a/2.. for times after 2t... The example illustrates the general feature of two-dimensional schemes that there is some initial preparation of the system [e.g. the X t (7f!2) pulse], then the system evolves for a lime interval tl under one effective Hamiltonian (in our case, one in which the spin-spin couplings are zero), then data are collected during a second time interval, t2. for which the Hamihonian lakes on some new effective Conn (in our case the full Hamillonian). 1lle experiment must then be repeated for other values of tl_ These three intervals are often referred 10 as the preparation, the evolution, and the acquisition intervals, respectively. The experiment we have just described is in (act essentially the famous experiment invented by Maller et al. [7.60] to resolve the C l3 chemical shifts and C I3 _H I spin-spin couplings. They demonstrated the technique using n-hexane (Figs. 7.35 and 36). They utilized broad-band (noise) decoupling of the protons instead of applying a "" pulse as in our example, and they decoupled during the second time inlerval. t2. instead of during the first time inlerval, tl' as we did in our hypothetical example_ They point out one could choose to decouple during either interval, and remark that since each point tl requires a separate experiment, one needs less data if the tl spectra is chosen to have the fewer spectral lines, Le. is the time during which the protons are decoupled. In a later paper, they describe the sequence of our example [7.61]' This general method is referred to as J-resolved 20 NMR in the literature. Qoli"CH;Qol;0i;CHrQol3 abccba
'"
.,
H'
". 0'.0
I,
I,
Fig. 7.35. Schematic representation or one ronn of C 13 2D-re;olved Spedroaoopy aner Miller ct al. 17.601. The CI3 nuclei are the I-spins, Iii thfJ S-spins. For the inlerval II the fullUamiltonian acts, but during t, the spin-spin interaction i, lUnled ofT by broad-band decoupling. The situation is similar to S-f1ip-only resonance with the role of the times t l lind t2 inlerdlllllged
322
Fig. 7.36. DlIla of Muller, KUI1t(Jr, and Ernsf using lhe pulse !lCquence of Fig. 7.35 for the CI3 spectrum of n-hexane (Clb-CII,-CII,-CII,-CII,-Clb). The ""I-axis gives the combined chemical shin and spin-spin spliuing, the loo'2-axiS the chemical llhirL5. Thesoe data show how a 20 display enables one to lIl""atly simplify unraveling a complex spectrom. TwentylW() experimenL5 Vooe"" coadded for each of the 64 values of tl between 0 and 35 IllS. The authors state lhat the resolution is seve~ly limited by the 64)(64 dab. matrill: used to represent thfJ 20 Fourier lransform. The absolute values or the Fourier coefficients a"" ploued. The undecoupled I D spe<;trum is indiCllted along the ""I-llll:is and lhe protondecoul)led spectrum is shown along the W2-all:is
323
I
'I
We shall discuss funher examples in later sections. First, however, we need to take a side trip to deal briefly with a topic the reader will quickly encounter in studying the literature, the problem of 20 NMR line shapes.
we can select either the real or the imaginary part of 9(Wl) and thus either the absorption signal or the dispersion signaL If, however, we take only the real part of g(WI,W2), it involves (for A real) (7.244)
7.23 Two-Dimensional FT Spectra - Line Shapes In the previous section we developed the Fourier transfonn of the theoretical expression (7.234) by fonnally extending the meaning of t1 and t2 to negative times. This is easy to do given a theoretical expression. If, however, we have experimental data, we cannot in general guess the theoretical expression. One then encounters a problem which Ihe reader will immediately come across if he reads the literature of 20 NMR. We therefore take it up briefly. A thorough treatment is given in the basic paper of Aue et al. [7.57]. Let us therefore reexamine how to treat experimental data considering positive times only. Since the problem is one of mathematics, we assume we have the simplest fonn of experimental data given by an exponentially decaying sine (cosine) wave. We put in the decay since such decays are always present, and to avoid difficulties at infinite times. Thus, suppose we have (7.239)
Then
1 9(Wt> W2)"" (211-)2
=
J =J dtl
o
i(OI
WI)J[a2
i(02
W2)]
O't+i(OI-WI) O'2+i(!t2-w2) = (211')2 + (0 1 - Wl)2 O'~ + (02 W2)2 A
(7.240.)
0
A
"" (211')2 [al
OI
(7.240b)
(7.24Oc)
In the corresponding one-dimensional case l(tl) = Ae(in\-aJ)l\
J
(7.241)
1= g(Wt) = 21l'
e-i"'lt\ l(tl)dtl
(7.242)
o we get
A 01 +i(JtI -WI) 211' + (01 - wd 2
( ) =gWI
O'I
(7.243)
We recognize that the real pan is the Lorentzian line shape (the absorption), and the imaginary part is the dispersion. If we use quadrature detection (Sect. 5.8) 324
A Re{g(wI,W2)}=-
al
2
0'20'1
(n
+ HI-WI
)2
(7.245)
which is the WI absorption line. However, the practical problem is that since dispersion extends out farther from the line-center Ihan does absorption, the mixing of dispersion tends to decrease the resolution between adjacent lines. There are various methods of processing data to deal wilh these maners. In addition to the treatment of Aue [7.57], the review articles by Freeman and Morris [7.56], Bax [7.58], and Turner [7.59] discuss these issues.
7.24 Formal Theoretical Apparatus I The Time Development of the Density Matrix
..
dt2/(ll,t2)e-llU\t\e-I"'2t2
1
which is a mixture of absorption and dispersion. Of course, if one goes along the frequency contour W2 - O2 , one has
In discussing the S-flip-only experiment (Sect. 7.19), where we were applying 'II' pulses to the S-spins, we assumed that the effect of the S-flips was simply to shift the precession frequency of the I-spins suddenly from one ms multiplet to the other. An S-spin originally pointing up now points down, and vice versa. If, however, one applies a '11'/2 pulse to the S-spins, classically they are now perpendicular to Ho. How then do we treat what happens in a quantum mechanical system? Classically, if Sz is zero, the S-spin will not shift the I-spin precession. That would lead to a precession midway between the two I-spin lines at WOI+a12 and WOl - a/2. But there is no transition al WOl among the eigenstates of the Hamiltonian, so this answer cannot be right. The next closest thing is for the spin to act as though it now has both frequencies. Thus, a spin which precessed at WOI + a/2 before the S-pulse would have components precessing at WOl + a/2 and WOl ~ aI2 after the S-pulse. As we shall see, this latter answer turns out to be correct. We clearly need a systematic and rigorous derivation of the result. Let us therefore treat carefully a concrete problem, a pulse sequence X 1(1I'!2)· .. r ... Xs«(J)···t-r
,
where we will explore and contrast what happens when (J "" 11' with the situation for (J := 11'/2. We wish to calculate Ihe expectation values of 17; and I y in the I-spin rotating frame. We shall use the doubly rotating frame Hamiltonian. 325
We assume that the X,(1C(2.) pulse is applied at t == O. At t == 0-, just before. we assume the spin system is in thennal equilibrium. (We also treat me case of 100 % polarizalion of the S·spins.) We employ the general fonnula (I+(t)}=Tr{I+en(t))
(7.246)
where 1+ refers to axes in the rotating frame. as does llR· Since we will be working almost exclusively in the rOlating frame, we drop the subscript R to simplify the nouttion. . Then, assuming 1{ is given by (7.162) except dunng the (Hlh and (HI)s pulses, we characterize states by m, and ms. getting
(1+(t»
=
L
(m/msII+e(t)lm/ms)
m"ms ==
L: (m,mslr+lm',ms)(m~mslelm[ms) "","'S "',''''s
(7.247)
Making use of the propcnies of the raising operator and of the eigenstates Im,ms). we have (m,mslr+lm/ms) ==
and -(fl[ + all). Therefore, no matter wfu:lt rotation angle we give the S-spins, these are the only frequencies we observe. At most we can change the complex coefficients
DmsmsDml,I/26m~,-1/2(,II+I-~)
(-mSIel + ms)exp (i(fl, - oms)ttl of the time-dependent tenns exp [- i(fl, - ams)t] Since the coefficients are complex, we have at our disposal the amplitude and phase of the two frequency components by manipulaling (-mslll(t\ >l +ms). To follow what happens in the actual pulse sequence, we use the method outlined in (5.167) in which we divide time into intervals during each of which the Hamiltonian can be taken as time independent. In SO doing, we will use the resulls of (5.253) and (5.254) for the time dependence of II under the action of rf pulses. We must first, however, generalize these equations which apply to one spin for the case of two spins. In so doing, we consider the effect on II of interactions in the rotating frame. Utilizing the concept that these interactions are large compared to the spin-spin coupling, we employ (7.206) and (7.2m) to treat the interactions one at a time:
:: Dms.m'sDm,,1/26m,,-1/2 Abbreviating +~ as +,
-! as -
(7.2530)
for m[, we thus get
1{s:::: -7S',(HI)SS~
(r(t)} = L(-msle(t)1 + ms) (7.248) ms Thus, there are only two matrix elements of e needed, one for each value of ms·
. During those times when me rf pulses are not turned on, we have, usmg (5.133), that
e
(7.250)
using the notation of (7.176). As a consequence
~(m,mslelm"ms)::*(m,mslll'HS : : r; L
'Hselm/ms)
(m,msle
mi',llI s
(7.254)
(m'lmsl1islmims) == 6m7m,(msl'Hslms) we get
(-msle
(7.251)
Thus, if we know II at time tt. we can express it for later times as amS)(t -
't)J .
!!.(m,msle(t)lml,ms):: dt
*L I
[(m/msle(t>lml,ms)(msIHslms)
m "s
-(msl1islmsXm[msle
This expression, even without specifying (-msle
(7.253b)
Since
11.:: -hfl,l~ -1iflsS~ + hollSl
(r(t» = L(-msll?<'I)! + mS)exp [- iUh m,
-1iwISS~
Thus, taking just Hs
(7.249)
where
==
(7.255)
Equation (7.255) shows that if we are driving the S-spins only (Le. 11, == 0, Hs ,;. 0), the I-spin quantum numbers do not change. There is therefore no mixing of matrix elements (m/msl/?(Olmims) with matrix elements such as (m,msle(t)lm'lms)' The problem is thus transformed to a one-spin problem. 327
As a result, we can immediately utilize the results of (5.253) and (5.254). Note, however, that though matrix elements of 7-{s are necessarily diagonal in ml, matrix elements of e are not. This is a very important point to which we will return later in discussing coherence transfer and multiple quantum coherence. Thus, under the action of 7-{s, acting for a time t, we can write (using ms for -ms) for the matrix element diagonal in ms:
(-mslg(tll +ms) = ![(-mslg(O)1 + ms) +(-mslg(O)1 +iiisl] +W-mslg(O)1 +ms) - (-mslg(O)1 + ms)] cos wIst i
+2:[(-mslg(O)1 + msl - (-msle(Oll + ms)J sin wlS'
(7.256)
We now let e
T
until time
T-
just before the S-spin pulse.
Utilizing (7.250) we get
(-msle(T-)[ + ms) = exp [- i(n , - amS)T](-msle(O+)1 + ms) =+iloexp[_i(n,_amS)T).
(7.261)
2 We now consider twO cases for the Xs(U) pulse, 8 ::: 7f and 8 ::: 7f12. Since (HI)S will not change the ml quantum numbers, and since (I+(t» arises only from teons ofT-diagonal in m/, we need consider only the effect of (HI)S on the matrix elements (-msle
and for the matrix element off diagonal in ms
(7.262)
(-mslg(tll + ms) = W-mslg(Oll + ms) +(-mslg(O)1 + msl]
For 8 = 7({l.
+!(-msle
+2:[(-mslg(O)1 +msl - (-msle(Oll + mslJ sin WIS"
(7.257)
An analogous pair of equations holds for the effect of (HI) I. Therefore, at t = 0- , just before the pulse X,(7f12) we have (mImsle
,
where 81 means the matrix at thennal equilibrium with the Janice tempernture. Utilizing (7.256) and (7.257), we see that (HI), will leave the matrix diagonal in ms, but will genernte elements off-diagonal in mI. Since (Wilt) = 7f12, we get at time 0+, just after the (HI), pulse, from the equation analogous to
(7.256l (-msle(O+)1 + ms) =
~[(+mslg(O-ll + ms) - (-mslg(O-)I- ms)]
which also implies
(-mslg(T+)1 + ms) = 1!<-mSIg(T-)1 +ms) + (-msle(r-)I +ms)]
isjust the net I-spin z-polarization at thennal equilibrium for those I-spins coupled 10 S-spins which have a particular ms. Since to a very good approximation ms is as likely 10 be as the (Iz(O-» is equally distributed between the two values of ms. Defining
+! -!,
(7.259l we get
(-msle(O )1 + msl =
(7.263b)
These results have simple meanings. For the 'II" pulse, (7.262) says that the 11' pulse interchanges the matrix elements of f! between ms and -ms· Thus. the matrix element develops under the action of ms up to T, ~en develops u~r the action of -ms. This result is just what we assumed m Sect.7.19, ~nd IS displayed in Fig. 7.18a,b, as well as Fig. 7.21. We will discuss the meaning of Denoting (I+(t» for a 8 pulse as (I-(t»o, we get for t>
(r+(t)). = ~
+
(7.263.)
(7.263) below. (7.258)
But
32.
(-mslg(T+ll + msl = W-msle(T-ll + msl + (-msle(T-ll + iiisl]
L: [(+mslg(O-)1 + msl -
T,
(-iiislg(O-ll- iiisl]
ms x exp [ _ i(n, - ams)r] exp [ - i(n , - ams)(t - T») .
(7.264l Utilizing (7.260), this becomes
(I+(t»ll"::: i~o "Ee-iI1,teiams(t-27')
(7.265)
ms
This result is exactly what we deduced in SecL 7.18 if one replaces T by the notation tll"' For t ::: 2T it displays the refocusing of the spin-spin coupling. ~e teon involving gives the chemical shift. lfwe set il, equal to zero and utlhze (7.260), (7.264) is identical to (7.171).
n,
iIo
2
(7.260)
329
For the case
e=1ffl we get
([+('»./2 =
~L
!«(+mslu
ms
x exp [ ~ i(lll - amS)T] + [(+insle1 + inS) - (-mslu
FIg. 7.33. The phllSe$ ~ of the I-spins wh)ch rault from application of a 11:/2 pulse to a 100% polarized S-spin sample (ms initially) at t = T. Bd'ore T, only one ~ is needed, corresponding to the single value or ms. The 11:/2 pullJe creates ill situation in which both ms = and _.1 arc present, hence leads to two ph_. Note that the same two precession frequcncies arc present as in Fig.7.37, but the amplitudes or the components differ corresponding to the different initial mS polarizations of the two e _
=1
•
+t
(7.266)
Again, utilizing (7.26), we get, for t > T,
(I + (t»
•
/2 =
iIo IonIt (e -e2
iflT
/2 +e- i '""/2) 2
x (e ia (L-T)/2 + e- iu (t-r)/2)
(7.267)
~is expression has a simple graphical interpretation (Fig. 7.37). The -(nIt) term gives ~e phase shifl arising from chemical shifts. The two terms exp(iatl2) and
exp (-Iai(l) show thai up umil time T there are two phases of magnetization, analogous to the ;1 and (/12 of Fig. 7.21. The product with the last bracket shows thai al t '" T. each of these phase Jines splits into two lines.
-(nl - a/2), the other precessing at -(nl + a/2) (Fig. 7.38). Thus the average phase angle advances at frequency -ill, as though there were zero coupling, but the effective lenglh of the average magnetization goes as cos(a(t - T)I2) and hence is not constanL The fact that the average phase is independent of a for the w:(1. pulse has as its classical analogy that the S-spins are pointing perpendicular to Ho, hence produce no shift in the I-spin precession. Indeed, this is the physical consequence of (7.263a) and (7.263b), which together show that even if only ms = is present after the rr(1. pulse, bolh ms :::: and ms = -~ are present after the XS(Tr!2) is applied. The sorts of considerations we have been exploring are treated in a different manner in two illuminating papers by Pegg et al. [7.62], and applied to various double resonance methods.
!
• Fig.7.37. The phllSe$ .p of the I-spins which re!lull from Ilppliulion of II. 11:/2 pul$C to unpolarized 5spir'lll at time T. Up until t = T, there are two phllSC$, one for each value of ms. At f = T the phllSe$ each split in two, but the t/opu are rela~ed so thu there are still only two pMCelSion frequencies for the 1. spins, although there are HOW four phllSe$
We can get added insight by taking the S-spins to be 100% polarized, so that ll[ t 0- only mS is to be found. Then, for t > T
=+i
=
(I+(t}}'/2 =
~[(++ Iu
(- + lu(O-ll - +)]
x exp [ - i(!h - a/2}T]
x (",PI-;(J)/-a!2)(t- T l];ex P[-i(J)/+a!2)('
T)]} (7.268l
This shows that up until t = T, we get precession at a single angular frequency, -(n l - an), corresponding to ms = but that after the 7:(1. pulse at time T, the precession splits into two components of equal amplitude. one precessing at
+!'
!
7.25 Coherence Transfer Now that we have employed fonnal methods to follow in detail the role of w:(1. or pulses, we return to one of the important applications of spin-coherence double resonance, the transfer of the magnetization of one set of spins to the other. The method of doing this transfer was discovered by Feher in 1956 (7.63,64]. In 1962, Baker discovered a related fonn of nuclear-nuclear double resonance which he called INDOR. Baker pointcd out that it was much like Feher's ENDOR experiment [7.65]. In 1973, Pachter and Wessels [7.66] invented a pulsed version of the INDOR experiment, which they named selective population inversion. leener's proposal {7.55] for two-dimensional NMR, as we shall see in Sect. 7.27 includes among other things a homonuclear pulsed version of the Feher experiment. Then a heteronuclear pulsed version, with all the benefits of Fourier transfonn, was proposed by MaJuisley and Ernst [7.67], with modifications by Bodenhausen and Fneman [7.68], Maudsley et al. [7.69J, and Morris and FnelTliln (7.70).
'If
3JO 331
Feher was studying the electron spin resonance of donors in Si doped with phosphorous. His work was part of the research that led also to his disc:overy of ENDOR (Sect. 7.7). He showed that he could transfer the electron polarization to the p31 nuclei. Since his experiment is conceptually very simple and straightforward, we begin with it. In keeping with the model system of Sect. 7.19, we assume the energy levels are those of Fig. 7.39a labeled by jm/ms), and also labeled 1,2,3,4 as shown. In thermal equilibrium, the populations of each state are given by the Boltzmann factors, which, in the high temperature approximation, may be wriuen as
p(m/,ms) =
~
[I
+
(nWOlm/ :;iwosms)]
(7.269)
where P(mb ms) is the probability of finding a given 1-$ pair in the state Imf, m~). In this ex.pression we have neglected the small spin-spin coupling in compUt1ng ~he .ener~es of the levels. The leading term, does not pr<Xiuc:e any net magnebzabo~ Sl~nals, so it is only the terms involving m/ and ms which express net polanzallons.
t,
la'
l_._
1
fbI 4
-i-
1.-0.
---i°.1
0
oA - 6
6
Fig. 7.39. The energy levels (a) and therm.al equilibrium population I)robabilities (b) of the model system. States are labeled accordlllg to the convention lm/ms)
L:
(Iz) = =
Lml [LP(ml.m s)]
hwOI
8kT ;: Ll
(7.270.)
!twos
8kT ;:,
L ms [LP(ml. ms)]
ms
t + (.1 + 6) p(-+):::: t - (.1 - 6) p(++)::::
Slate 3 P(+-}=1+(Ll-6) 332
(7.273)
III/
Now, though we have neglected the spin-spin coupling in computing the thermal equilibrium populations, we must tIOt neglect them in calculating the spectrum since they make transition 1-2 have a different energy from that of transilion 3-4. Since the" in (7 .27Ia) is the same in all energy levels, it has no effect on the intensity of any transition, and can thus be conveniently dropped in recording population probabilities, giving one the population excess above (A negative excess is a deficiency.) These are conveniently expressed in a matrix
t"
i.
mS=l
ms=-!
m/
"'-!
- .1+S
- Ll-,
m/
"'+~
Ll +'
Ll- ,
(7.274)
Thus, the resonances for both I-spins and S-spins are doublets, as shown in Fig. 7.40. For the sake of explanation, we assume the I-spins have a much larger population difference (Fig. 7.403) than the S-spins (.1:> S). So that we see a larger intensity in the I-spin absorption lines (Fig.7.40b). Feher's method consists then of two steps. In the first, he makes an adiabatic passage of frequency w/ across one of the I-spin transitions (but not the other) such as the transition 1-2 for ms '" Since the adiabatic passage invens the magnetization, it interchanges the populations of the states involved, I and 2. This step leads to the population excesses shown in Fig. 7.4la, and listed in the following matrix: ms'" ms"'!
-!
- Ll-' Ll- ,
(7.270b)
we have for p(m/,ms)
Slate 2
(7.272)
ms
ml
(S,) =
(m/msIIzlm/ms)/l(m/ms)
+!.
Defining
State I
(7.271d)
The spin polarizations are
t -A·6 1-"'06
fmjmsJ
P(--)=t-(.1+6)
State 4
(7.271.) (7.271b) (7.2710)
(7.275)
The S-spin population difference corresponding 10 a tranSition in which ms changes, such as fTOm 1- +) to 1- -) (states 2 (04), would be given by
P(-+)-P(--)=(1 -Ll+6)-<1 -Ll-6)=20 bl!fo~
(7.276.)
the adiabatic passage, but 333
Energy
Energy
F1g..':40: (a) The energy level and lhcrmal eqUlhbrlUm populations (sc.henlalic) for a CIlS(! in ~hich "'II" "'Is !lID that 11" 6. (b) The I-spon and 5-spin resonance intensities sho,,:,in~ the ~pin-spin splitting. The grealer I-spm lIItell.'llty an!leS bec:aUS1e "'If:> "'Is
la) 1- .J
""/---
r---""
+i
P(++) - P(+-) = 26
(7.r17.)
P(++) - P(+-) = -2il- 26
Population
I-Spin Absorption
S-Spin Absorption
Energy
lal
(7.277b)
which, for Ll"~, is an invened population. The population differences for both (7.276b) and (7.277b) are much bigger after this first step than before. Since (7.277b) corresponds to a population inversion of the S-spins [the upper energy state ImJms = 1+ -) has a larger population than the lower energy state ImJffis) = 1+ -)], there is no net polarization of the S-spins. To achieve a net polarization, one may therefore inven the population difference of eidlCr (7.276b) or (7.277b). Feher achieved this by his second step, an adiabatic passage of (H 1)5 across either (but not both) of the S-spin doublets. Thus, if the passage were across the 1-3 transitions, the 1-3 populations would be interchanged leading to
"s; 112 .s: -112
Energy
-!,
after the adiabatic passage. Thus, for ml = the S-spins have an enhanced polariz,ation. On the other hand, for the S-spin transition involving ffil = (the transition between states I and 3), we have a population difference before of
but after of
Population
Ib)
(7.276b)
p(-+) - P(--) =(t + il +6) - (t - il- 6) =2il+26
1--1
"'5= , "'5=-'
""1---
}---,.-,
il + 6
- il- 6
il- 6
- Ll +0
(7.278)
Using these populations, one gets readily
(5,) = !(il+6 - (-il-6)+(il-6) - (-il + 6)] = 2il
(7.279)
compared to its value before the two Steps Population
Population
(5,) =26 t-SjNl Absorption
S-Spin Absorption
Ibl
ml "1I2 ml :-1I2 li. 7,.41. (a) The ~ergy Jevel population. afler an adiabatic '--a•• 'hroo.h ,h. 1 . ,~me or rns = +.L .how· lh ,.. ..,..-spon level hassmal/u l ' I .1IIS e popu.atlon lIIverslon III which the lower energy 1+ +) popo atJ:On lh~n the hIgher eners,y +) level. (b) The J- in a d 5- . res
ms:1I2 "s·-1I2
1-
.
(7.280)
Such a scheme of magnetization transfer can be used for transferring electron (I-spin) polarization to nuclei (S-spins), or for transferring large "'( nuclear magnetization (protons) to lower..., nuclei (e.g. C 13 ). In order to be able 10 carry out an adiabatic inversion of one multiplet, one must be able to pass through the multiplet with an HI sufficiently small to inven one but not the other lTansition. Thus "(HI must be less than a. AnOlher requirement is that spin-lattice or spin-spin relaxation times not prevent the adiabatic condition in this time interval. Thus, since it takes many precession periods I/"'(H1 to have the passage adiabatic, we have ,H) :> lIT" 11T2. The Feher approach to coherence transfer using adiabatic passage can be replaced by a pulse technique, the method used by JUller [7.55J and by Maudsley and Ernst [7.67]. Although superficially (7.279) and (7.280) suggest that one is
334
335
performing an Overhauser effect, in fact, since the method does not use spinlattice relaxation, it is clearly based on different principles. We now show how to use pulses to invert one hyperfine line. Consider an WI tuned exactly to the unsplit I-spin resonance frequency Wo/. (7.281)
WI = WOI
AI t = 0 we apply an XI('If(l.) pulse (Fig.7.42a). The magnetization vectors of the mS and mS lines are shown in Fig.7.42b at t 0+. We label them MI(~) and MI(-!) corresponding to ms = and -~ respectively. They now precess in the left-handed sense at angular frequencies (WOI - amS) in the lab frame, or at -oms in the rotating frame. (This is equivalent to precession in the rotating frame frequency oms in the right-handed sense, see Fig.7.42c.) At a time, to, such that
=!
=-4
=
+!
oto = 1f
(7.282)
they are pointed opposite to one another as shown (Fig.7.42d) along the x-axis. At this time (Fig.7.42e), we apply a 'fro. pulse with (HI)! along the y-axis Ia Y/(1fn) pulse]. This puts the M I (+!) along the negative z-axis, and MI(-!) along. the positive z-axis (Fig. 7.42f). Thus, we have achieved a population inline, while leaving the ms -~ line with a normal version of the ms population difference. This is equivalent to an adiabatic passage across just t~e ms = line. There are two important points to n~te. The ~rst .is that t~e sphtting, 0, is essential since il is responsible for prodUCing the suuatlOn of Fig. 7.42d in which the two components of AI I are pointing opposite to one another. The second point is that the phase of the (H 1)1 pulses (Le. their orientation in the x-y plane) is important. If, for example, the second pulse had been another XI(1fo.), it would have been parallel to the MI vectors, and would have had no effect on their orientation. Likewise an X('lfo.), where X means (H')I pointing along the negative x-axis, could be used to obtain a state in which the MI(-!) would be inverted instead of MI(!). So far we have only inverted the ms = transition, therefore we have achieved the effect of the first of Feher's twO steps. To complete the coherence transfer, we must now do a similar thing for the S transitions_ Thus, with ws tuned to Wos, we can apply the sequence XS(1fo.) followed at a rime delay to = 'If/a by a YS(1fn) pulse to invert the Af, = transition for the S-spins. This would have completed Feher's coherence transfer and produced the populations in the matrix of (J .278). If instead of applying four pulses we apply the sequence
=+!
=
!
!
lal
L [H,.II
lei 11,11/21
Ibl~
Hl ll121
MI I-1I21
+i
"
-I
"
Idl
, ,,
"
t1 1 11/21
~
" (fl
"
(a)
Mil-liZ!
[Hll l
"
"
Fig.7.428-1. The effect of (lid! pulses tuned to resonance (WI = ,",od on the I-spin magnetization M,(ms) for the two components ms = and ms _1. (a) At I 0, (Ildl is applied IIlong the z-lIxis in the rotRting frame giving an x/('II'/h pulse; (b) I'll I = 0+, both At I (t) and M l( lie along the y-axis; (c) the two components M /(mf) precellS in o[.posite diredions at angular frequencies ±a/2; (d) at a time t 11:/0, 111'(2) and MI( - ~) point along the -z- and +z-axes respectively; (e) an (HI)I lying along the y-axis gives a Yl(lr!2) putse; (f) the I-magnetization vcdors after the Yt(,../2), with MI(-i) pointing parallel to the ~-dire<:tion, but MICt> pointing antiparatlel
+t
336
L (H,ls
1'1 1 11/2)
-!)
•
(7.283)
the I pulses will have produced a situation in which the M I (!) population is inverted, bUI not Ihe MI(-!). so the situation will look like Fig. 7.4Ia. Therefore the XS(1f(l.) pulse (Fig.7.43a) will produce a silUation in the rotating frame such
1'1[1- 112 1
" lei
X,(.12) ... (./a) ... (y/(./2). Xs(.12)) .. . observe S
/11.1-1121
=
=
Ibl
" Mi l-1I21
" Ms !1I21
(e I
=
"
"
~·Ig. 7.43. (a) (/fils is applied along the :I:-dire<:tion in the S-llpin rotating frame, giving an Xs(rr/2) pulse; (b) the SlIpin magnetization Ms(m,) jUlIt after the Xs('7f!2) putlle. Ms(_l) and Ms(+l) are antiparaUel; (c::) /lS a result of their dMerent precell6?on frequencies, M s( ~) and M s( -~) are parallel after II. time 1t!IJ
337
as shown in Fig. 7.43b in which the two components of Ms point opposite to one another initially. At later times, their relative phase will have changed, making the magnetization components, in fact, parallel to each other at time 1r/a after the XS(7f/2) pulse (Fig.7.43c). Indeed, following the XS(1f/2) pulse, there will be an S·spin free induction signal. In the lab frame, it can be thought of as two signals, one at wos - 012, the other at wos + 012, or it can be viewed as a signal at wos which is amplitude modulated as sin al/2. This signal is of course similar to what would result from using Feher's two adiabatic passages to transfer the I polarization to the S-spins, followed by a 1r/2 pulse to the S-spins to generate magnetization perpendicular to Ho. If one performed that total sequence using pulses instead of adiabatic passage, it would take a total of five 1f/l pulses (two of the I-spins, three for the S-spins). However, the last two "/fa. pulses can be omitted since the first of them rotates transverse S-magnetization to lie along Ihe z-direclion, and Ihe second rolates the magnetization from the z-direction back to a transverse orientation. In their technique of selective population transfer, Pach/er and Wessels [7.66] apply a 1f pulse to one of the I-spin transilions, invening their population difference. They use an (Hth much smaller than the spin-spin splitting. Then they apply a strong 7f/l pulse to the S-spins 10 observe the free induClion decay. The Fourier transform of the S-spin signal shows one hyperfine line invened with respect to Ihe other. and a large signal increase if "'(, »"'(5' Thus their experiment is equivalent to (7.283). For the sequence (7_283) we have required that = wo" exaCI resonance for lhe I-spins. For Ihe S-spins, XS(1fn) does nOI have 10 be exactly at resonance. It will still produce the situation of Fig. 7.43b. Then, those magnetization components will precess freely at (wos - am,), giving a signal at WO/ which is amplitude modulated at angular frequency o. The sequence (7.283) is almost the Maudsley-Emst sequence. What they do is (i) remove the requirement that WI = WOI and thereby (ii) remove the special phase requirement of the second I pulse. Then, taking a sequence
w,
(7.284)
in which t2 is the observation time. and for which (7rn), means an I-spin 1f/l pulse without a special phase condition. they make a 20 Fourier transform. The fact that WI "WOI produces a situation in which. following the initial (7f/l)/> the I-spin components precess both with respect to one another and also with respect to the rotating frame. Thus, if one is off resonance by an amount large compared to the spin-spin splitting, the pattern of Fig. 7.42f will rotate rapidly at (WOI -wI) in the frame while the relative orientation of M/(!) and M,( will change slowly. The situation is shown in Fig.7.44. The initial X,(7fn) (Fig.7.44a) puts M,C!) and M/(-!) along the y-axis in the frame rotating at (Fig.7.44b). Under the combined influence of chemical shift (il/) nnd spinspin spliltings the M, vectors rotate and separate (Fig.7.44c). A pallem of the }.th's at successive times is shown in Fig.7.44d-h. If another 1r/2 pulse with
w,
w,
338
-!)
101
L IHII I
Ml l1121 M11-1I2)
"
" r,
Idl
y, IOI-al2)1
1'1
.
Ibl
M111I2)
to l • anJt
" lei
t-
............ ~t-V21
y,
If J
MI II121
,1 M 1-1/lJ ,1
MI III2I
,
" I
I
/
Ihl
y,
191
"
,,M I-1I2)
"
I
y,
M11-l/l1 MI I-ll2)
----
-
Ml l1l2l
'.
" M1l1/l1
T
=
Fi&o 7.44a-b. The effect of the. '_spins bei~g off resonance ("" 1+'0,). (a) At t 0, (lid, is applied along the lI;.axis m the rotating frame; (b) the .)1.,(_/2). pulse puts th\t'"':O componenlll of MI(ms) along the y-axis at I = 0+. M/(t) IS the solid arrow, M/( - 2) IS the dashed arrow, (c) the componenls precess at 01 -: am~, where.OI = ""01.- "'I, un~er the combined effect of the chemical shif~ (flt) and IIpm·spm coupling (ams), (dl--:: compared to their joint rotation rate 01' An XI(_/2) pul!le would have no effect (or sltu~tlOn 00, wo~ld make M I( 1) inverted rdative to 110 for situation (e), and make M I( IIlverted relative to 110 (or ~tuation (h)
hI
-!)
(H t) J along the x-axis is applied (Fig. 7..44a). ~ere will be comple~e in.version of M,(!) for the situation (e), complete IflVerslon of M,(~!).r0r sl~unuon (h), no inversion of either component for situation (g), and partial II1verslon of ~h and M (_') for situations (d) and (t). We see therefore that the chemical M I ( "',) 1 /"'1 . ' shift frequency flj will determine the urnes at whIch the t':'"o M, com.ponents are invened, and thus they will determine the times til which a~ S-SPIfl pulse will find transferred magnetization. Of course, a will also determine the general 339
range of times for which Mj(!) and Mj(~4) point opposite to one anOlher. Indeed, Bodenhausen and FreelrUJn [7.68J analyze just this model to show that if the I-spins are protons and the S-spins are C l3 , the resulting 20 spectrum has the proton spectrum in the WI direction and the e 13 spectrum in the W2 direction. Thus by observing the C I3 spectrum one gets the indirect detection of the proton resonance frequencies, and the correlation of the proton and carbon chemical shifts. It is useful to think of this experiment in tenns of the density malTix analysis of Sect. 7.24. We have a pulse sequence Xj(1f!2) ... r ... (X j (1f!2), XS(1f!2» . .. observe S
(7.285)
in which we shall treat the X S(1f!2) as having occurred just after the second Xj(7f!2) pulse. Using our theorem of (7.207) we define r- as just before and r+ as just after the second X j (1f!2) pulse, r++ as just after the XS(1f!2) pulse. Since eventually we will observe the S-spins, we seek (mj-jg(t)lmj+) during the observation time t>r++. Now (mj_lg(r++)lml+) is prcx1uced by XS(7C!2), hence from elements of e{r+) which are diagonal in mi. At t = 0-, before the first Xj(wll) pulse, we assume f! is in thennal equilibrium, hence diagonal in both ml and ms. Hence, under the action of the two X j (1f!2) pulses, f! remains diagonal in ms· Thus, (ml-Ie(r++)lmj+) arises solely from elements of e{r+) just before the pulse XS(7C!2) which are diagonal in both ml and ms. These mamx elements can be traced back to elements (mlmsle(r-)Imlms) as well as to elements (=F msle{r-)l± ms) using (5.253):
(mjmsle
4[(mlmsle<0-)lmlms) + (mpnsle
.. + 4[(mjmsle(0-)lmlms) - (mlmsle
for the diagonal elements just before the XS(7C!2) pulse. There are several interesting limiting cases. First, if n j = 0, the result depends on cos (amsr), hence is the same for ms = and ms = This result confinns our poinl earlier thai in the absence of a chemical shift an Xj(7r!2) ... T ••• X j (1f!2) will not inven only one multiplet Instead, one needs an Xl(w!2) ... r ... Yj (1f/2). In fact, if f1[r = (7r/2), one has the case of Fig.7.44g, and indeed then and cos (ill - amS)r = sin (amsr) so the result differs between ms = ms = ~!. Thus one can inven one multiplet only. Note also that as T goes to zero, (mlmsle(r+)!mlms) goes to (mlmsle<0-)Imlms), representing the fact that two closely spaced 'I({l. pulses are equivalent to a single 7f pulse. We can now use (5.254) to get the S-spin magnetization
4
(S+(t,» = T, (s+ e
m,
=
(7.286)
and
(Tmsl,(O+)I± ms) = ± ~1(+msle
and
(7.288)
2
m,
(7.289)
(mf - le
(7.291c)
where the malTix elements of e(r+) are given by (7.290). Equation (7.29Ic) is the mathematical expression of the fact that the S-spin signal is detennined and ms = for by the population differences between the states ms = each value of ml' existing just prior to the XS(7r{l.) pulse. These population differences oscillate as a function of T, the time difference between the Xj(1r/2) pulses, at the nltes [} I - a/2 and ill + a/2. We can utilize (7.290) together with (7.275) to follow what the two X I (Tr!2) do to the population excesses. Thus we have for t = 0- , the population excesses
!
ms
We can thus relate the diagonal elements of f!(r+) to the elements of e
~ L: [emf + le
(±msl,(O+ll±msl = t!<+msl,(O-)1 + ms) + (-msle
(7.29Ib)
m,
;
(7.287)
(7.29la)
1e<',}lmf+)
= L(ml _le{T++)lml+)e-icns-amt)t2
± 1:[(-msle
-4.
+!
(± msle
- (+msle
(7·290l
=1
ms
=-,
ml=-!
-Ll-o
-Ll-o
=!
L1+6
Ll-S
ml
-4,
(7.292)
which become, for t = T+, just after the second I-spin pulse, using (7.290), 341
ms=,
-i m/ = ! m, '"
ms=-!
6 + Ll cos (ll[ - aI2)r
- 5 + L1 cos (01 + a!2)r
6 - Ll cos (II{ - 012),)
- 5 - L1 cos (n l + a!2)r
Therefore, from the point of view of the S-spins. the m/ ::
difference (+ +
31
M/(ms) =
!
and since ..
(7.293) line has a population
T[(+msle<'+lI + ms) - (-msle<'+)!- ms)J .
(7.297)
(+msle<'+)1 +ms) - (-msle
"" 26 - A.[ cos (0/ - aI2)T - cos Uh + oI2)r]
(7.294.)
-i has correspondingly
= [(+msle
Moreover,
(+msle
(- + 1e<'+)I- +) -
(- -1e<'+)I- -) '" 26 + L1[ cos (ill - a/2)r - cos
(7.294b)
Bodenhamer! and FreemtJn derive the resuil of (7.290) by use of the classical spin picture (Fig. 7.45). The X t (7Cn,> pulse (Fig. 7.45a) puts the [-magnetization along the y-direction in the 1 Jl)(ating frame (Fig. 7.45b). This precesses at (0/ams) (Fig.7.45c). The second X,(7fn) pulse r()(ates this vector down into the z-z plane, with a .:-component M,/(ms) given by {Fig. 7.45<1)
(7.295) This equation can be converted to an equation for elements of the density matrix
since
r, I: O'
(bl
(7.299)
Combining (7.298) and (7.299) we get (7.290). In summary, the pulses work by having the X pulses take (m/msle(O-)lm/ms) into (±msle
l
L
(e)
(7.296)
Therefore, using (7.295)
1e<'+)1 ++) - (+ -1e<'+)1 +-)
('I
ms)]
"( h
M./(ms) =
t '" r+ of
and the m/ ::
1~r, [(+msle
Yl
t •
y-
Idl
.,
"
(7.300)
r,
would refocus both the field inhomogeneity and the spin-spin splitting. However, thinking back to spin-echo double resonance, we see that if we also flip the 5spin by a 11" pulse simuhaneously with the [-spins, the spin-spin coupling will not refocus (see discussion at the end of Sect. 7.21). Thus, the sequence HI (lIl s 1 (OS (1'1 1 -
im s) 1
=
Flg.7.4Sa-d. The efroct of two X,(1f/2) pulses, one at t = 0, the other at I T on the I-spin magnetization M I (illS). (a) The diroctioJl of (Jh), for the X I (1f/2) pulses' (b) the magnetiulion M/(llIs) lies along the y-axis all 0+; (el Mdms) processes thr~ugh an angle (nl - (lmS)T at time T-; (d) an X 1(>2) pulse at I T puLs /If1(ms) int(l the .7:-~ plane, with a eoml>onent along the I-axil! of -Al t cos (n, - oms)T
=
342
=
x/(,m ... , . .. (X/(,), Xs(,»···,··· y/(,m
(7.301)
2ra =
(7.302)
with 11"
will produce a shuationjust after the last pulse in which the M/(!) magnetization is inverted, but not the M/(-!) magnetization. 343
Then, a 7r/2 pulse applied to the S~spins will generate transverse magneli~ z.ation. It, too, will dcphase if there is magnetic field inhomogeneity, but can be refocused by an X S (7r) pulse accompanied by an X/(;rr) to prevent the spin-spin spliuing from being refocused. The same condition 2Ta = 7r of (7.287) will guar~ antee that at 2T following the first Xs(-;rr/2) pulse the magnetiz.alion will look like Fig.7.43c. We can write this sequence in a slightly different format as x/(7fn) ... T
•••
X/(1r) ... T
•••
Y/('Il"I2)
X/(7f)
i
XS(1fI2).··T",Xs ('II")··.T ... S-echo
Xs('II")
(7.3OJ)
with
+
(I (t» (7.304)
2aT = 7f
in which we utilized two lines. one for the I-spins, another for the S-spins to help keep track of the two spin systems. If the "inhomogeneity" comes instead from the existence of a number of chemical shifts of the I-spins, then one can still transfer I-spin magnetization to enhance S·spin signals: X,(7fI2) ... T
•••
X,(7f) ... T
XS(7f)
•••
Y,(7fn)
X S (1rI2)···acquire
(7.305)
in which we assume the values of a are rather similar for all chemical shifls. This sequence was discovered by Morris and Freeman [7.70] who gave it a name "INEPT' to stand for "insensitive nuclei enhanced by polarization transfer".
7.26 Formal Theoretical Apparatus lIThe Product Operator Method In our example of coherence transfer, we saw that it was important to have a formal method to carry out a calculation in order to be sure one was dealing correctly with such things as the effect on one spin (e.g. the I-spin) of applying a 7r/2 pulse to the other (5) spin to which it is coupled. The differential equations for the density matrix gave us such a method. However. the density matrix for two spins (m/msll?lm',lns) has sixteen elements, since m" ms, In J and Ins can each take on two values. 'l1lerefore, the differential equations are really quite complicated. We picked an example in which only a few elements were involved and in which we could see clearly just which elements we needed. Naturally one would like a formal method for calculating the time development of the density matrix which is straightforward to apply, and which is readily extendable to cases of more than two spins.
344
We tum now to such a method. which has been named the "product operator" fonnalism. This method was developed independently by at least three different groups: Sorensell et al. [7.71], Van den Ven and Hi/bers [7.72] and Wang and Stichter [7.53.73]. We shall present the method in connection with the two-spin system. The generalization to more spins is straightforward [7.71]. We confine ourselves to spin nuclei. The typical situation one encounters is illustrated by (7.212) which describes spin echo double resonance:
=
+
11WO,
.
ZkT Tr{I [exp(l(il/I:: + ilsS:: - uI:5::)(t - T»X/(l'l")Xs(7f)
x exp (i(il/ I, + ilsS:: - aI:I:)T)X,(7rn)I::xil(7rn) x exp(-i(il,I:: + ilsS, - aI:S,)T) (7.212) l x X '(7f)Xi ('Il")exp (-i(il,I: + ilsS: - aI:S:)(t - '7f»)}
s
The system has an initial density matrix corresponding to thennal equilibrium given by (7.210): JO-) = ~
'Z!'(t +
hwo/I: +IIWOS
IT
S:)
a21~
The effect of the rf pulses is to produce rotations. We may symbolize these in general by unitary operators R. An example of such an R is X/(9): X/(9) = ei 'z8
(7.306a)
Xil(9) = e- i 'z8
(7.306b)
Likewise, between pulses the system evolves under the influence of the time development operator T(t) given by (7.307.) a·J07b)
(7.308.) a.J08b)
(7.30&) Since T/. Ts. and T/s commute with one another, they can be applied in any order in (7.307b). We have already worked a good deal with the R's and T's. We use the symbol U to stand for R or T and U-I for their inverse. Then if Ii (I, S) and !2(I, S) are two functions of the spin components of I and S, 345
uh (I, S)!z(I, SW- I = UII (I, SjU- 1U f2(I, SW- I
(7.309)
Hence, expanding the exponential in a power series, applying (7.309) term by lenn, and converting back to an exponential, we gel Uei!J(J,S)U- =eiu/J(I,s)U-l 1
Since the same is true for in (7.315), obtaining
cos (a5 z t l)
= I _
si or S;, we can replace S; by i
(at~~)2 + (at~~)4
[51 by (-t)2, etc.] (7.319)
.
(7.310)
Now, the R's plus TJ and Ts are all simple rotation operators of the general form exp(iI:I;8), exp(i/yO), exp(i!..,O). Their effect on a density matrix e(O-) such as (7.210) can be easily followed using equations such as
cos (aSztl) = cos (atd2) For (7.316), we get d3 sin (as,'I) = S ( , _ (at + (at5!1)5 S4z + ... ) z a 1 3! s2z
(7.311) based on (2.55). We can constnlci fannulas such as (7.311) by simply noting that when the exponential on the left has a positive exponent it generates a left-handed rotarian () of the vector, I;:, sandwiched between the two exponentials. The problem arises from TIS in expressions such as T/s(t\)IyT/- I (tt) = e-illJ.S.IllyCiaf.S./l S
(7.313) then (7.312) looks much like (7.311):
(at
l _
z
2
(7.315)
3!
...
)
(7.320)
sin (aSztl) = 2S z sin (atd2)
(7.321)
Therefore, utilizing (7.319) and (7.321) in (7.314), we get e-ialzS,tlIyeialzS,tl = I y cos (atd2) - I;r.S z 2 sin (atd2)
(7.322)
It is useful in carrying out calculations to have these various results in a table (Table 7.1) in which we list U It;rU- 1 for
1.
(7.314)
This equation has an operator, Sz, as the argument of the cosine and sine, a situation which is quite acceptable if we recall thal for operatOrs the trigonometric and exponential functions are defined in terms of their power series expansions. It turns out we can simplify these expressions by starting with their power series:
(a~~)3 (~y + (a~;)5 (~y + ... ]
= 2S (atl) _ (atd2)3 +
(7.312)
which have bilinear forms (lzSz) rather than linear forms (Iz) in the exponent. If we now consider
e-ialzS,tIIyeiCllzSzLI = I y cos (aSztl) - I;r. sin (aSztl)
= Sz
Table 7.1. The effect of UloU-l for various U's and for a = "',y,z. The bottom line applies for spin only
t
I,
I.
I.
l~cosO-l.sinO
lrzcosO+lzsinO lrz cos 0 - l~ sin 0 I" cos(Oj2)
I,
I z cos 0+ I~ sin 0 I, cos 0 - I" sin 0 I. I.
I.
I.
e i9frz ei9f~
U
e i9fz e- l9fzS •
+1~(2S, sin(Oj2))
I~ I~
cos 0 + Irz sin 0 cos(Oj2)
-lrz (2S z sin(Oj2»
(7.316) We now make use of the spin! property. If we choose as a basis set functions lms), (mS = or the most general function is of the form
+!
1{;(t)
-!),
= C+(t)lms =!) + C_(t)lms = -4)
(7.317)
Explicit evaluation gives S;1{;(t) = C+(t)S;I~) + C_(t)S~1 -
!)
= iC+(t)I~) + iC-(t)1 -~) =
346
i,p(t)
We now see how to evaluate TISIt;rT1S1 for a = x, y, z. An interesling point is that it takes single spin operators, It;r, into a sum of a single spin operatOr and a product of spins It;rSjJ. We see therefore that if applied several times, it might develop terms like It;rJjJS.rS6' But such terms can be simplified as we now explain. Suppose we think of either the I-spins or the S-spins in terms of the 2 X 2 Pauli matrix representation. We recognize that spin products such as I",Iy are products of one 2 X 2 matrix with another. The result is also a 2 X 2 matrix.
(7.318) 347
Now any 2 X 2 unitary matrix can be represented as a linear combination of the three Pauli matrices plus the 2 X 2 identify matrix. Thus. it must be possible to reduce products such as Izly to linear combinations of the identity operator with la's. In fact, we have shown one such example of a product Ialp for the case that cr ::: p:
.
1~:::1
cr:::x.y.z .
We shall show that if operator. The result is iI,
2
Izly :::
Q
(7.323)
:I P. we can also easily reduce Ialp to a single spin
111 1z :::
B1 :::at,
iI,
-2
(7.324)
(7.325)
This relationship, however, also involves I y l z • so we need one more relationship involving l z I1/ and I1/Iz . Anolher such relationship arises from products such as (1z +1,)2. Let us then consider the coordinate transfonnalion from (1z,I,) to (1z" III) wilh I
I., = j2U, + I,)
(7.326)
a 7r/4 rOlation. Then (7.327)
i ::: !(1; + I;) + ~(1z1y + 1,lz) ::: !
l z 1y + lylz ::: 0
t, ... XS(7r{2) ... t2
Then
r~1-1 e
X S I (,...fl)T/f}B2 .
(7.331)
The effect of the XI(,...fl) pulse is to rotate I, into l y : XI(1rfl)1,X/I(1rfl) = 111
(7.332)
.
Next we apply TIS' Utilizing Table 7.1 we gel TlS(B 1)Iy T ls)(9,) '" I1/ cos 9)12 - 21z 5, sin BI12
(7.333)
Next we apply Xs(,...fl), a left-handed ,...(1 rotation of spin S, giving XS(,...f2XIlI cos 9 1(1- 2Jz 5, sin 9 1(1)X I (,...f2)
S
= I y cos 9i1l - 2Jz 5 11 sin 9 1f2
(7.334)
We now act on this with TIS(92):
s
T/s(B2)ly T/ 1(B2) cos 9 1(1- 2T/S(92)lzT/-s)(92)TIS(92)5yT'SI(~)sin 8,(1 = (1y cos 9212 - 2lz 5, sin e./.fl) cos 8,(2
- 2(1z cos 82(1 + 2l11 5, sin 9212)(511 cos fhfl- 21,5z sin 82fl) sin 9 1(2 "" 111 cos 92fl cos Blfl- 21z 5, sin 82(1 cos 8 1(1 - 2Iz S11 cos 2 8212 sin 9 1fl + 41z 1,5z cos 82n sin 82n sin 81n - 41y 5,5y sin 82n cos 82n. sin 8,n. + 8Iy 5,I, 5 z sin 2 82fl sin B,n.. (7.335)
We now have a number of tenns involving spin products such as 1a lp or Sa5{J. Utilizing (7.324), we reduce the various teons as follows:
.
(7.336)
The initial density matrix at t::: 0 is given by (7.210). Suppose we consider the ponion of e(O-) given by
~~~I,
Evaluation of Ihe portion involving 5, is left as a homework problem. 348
(7.330)
.
(7.328)
for a spin ~ panicle. Substituting (7.328) into (7.325) yields (7.324). Let us now illustrate how Ihese fonnulas are used. Suppose we analyze the 5-fHp-only double resonance for the case of a 7r{2 5 pulse:
e(O-):::
B2:::at2
•
X
plus cyclic pennutations. To prove (7.324), we nOle thai tenns such as l z l l1 are encounlered in the commutation relation
Xr(7r{2) ...
For simplicity, we consider a case in which il r ::: ils::: 0 (exact resonance for both spins). We apply an X r (7f!2) pulse al t '" 0, and an Xs(Tr/2) pulse at time tl, let the system evolve under TIS from 0 to tl, and after t1 for a time t2. We define B, and fh by
(7.329)
l y 1,5,5:I: =
'2 5 ('2i) 17; (i)
I.S 11 "'-~
When these are substituted into (7.335), the two tenns in 111 57; cancel, and the teons in IzS y combine (cos 2 8212 + sin 2 92fl = 1) giving 349
TI S(8 2)XS(1r(2)TIS (8 1)X1(7r(2)I z Xii (7r(2)T1S1(8 1)X5 1(7r(2)Ti!] (82) = I" cos 82/2 cos 8 1/2 - I z S z 2 sin 82/2 cos 8 1/2 - I z S,,2 sin 8tf2
. (7.337)
Therefore, the contribution of (7.329) to
e is
e(tl + t2) = hWOI (lz cos 82/2 cos 9 1/2 - IzS z sin 92/2 cos 9 1/2
ZkT - I z S,,2 sin 9d2)
(7.338)
These various terms tell us what elements (mlmsl,q(t)lm~ms) are nonzero at time tl +t2' For example, the first term involves I z , hence gives nonvanishing elements (± mslu(tl
(7.339)
+ t2)I~ms)
The term IzS z gives similar elements, weighted with mS:
s
(mlmsllzS z Im~ms) = ms(±IIzl=f)Oms,m Omr,±Om/,:,:
(7.340)
Lastly IzS" givcs nonvanishing terms
(± ±!e(t, + ',li'!' '1')
(7.341,)
(7.342)
Tit"', + ',)1'1' ±)
(7.341b)
However, since two nuclei are involved, with II = IS but £2r i- £25 owing to their chemical shift difference, it is better to think of this as
(±
whose further significance we discuss in Sect. 9.1. In Table 7.2 we list the other useful relationships employed in dealing with spin operators. Table 7.2. Useful relations for unitary operators U. (U is a function of I, or 5, or I and 5)
U/J(1,S)h(I,S)U-l
U/l(1,S)U-IUh(l,S)U
I
Ucxp(i/(I,S)lU- 1 = exp[iU/(I, S)U-I I For spin
!:
cos (1.8) = cos (8/2) sin (1.8) = 2/. sin (8/2)
7.27 The Jeener Shift Correlation (COSY) Experiment Since thc early days of high resolution NMR, an important goal has been to tell which nuclear resonance lines arise from nuclei which are bonded to one another. The existence of bonding manifests itself in liquids through the indirect spin350
spin coupling [7.74,75] (Sect. 4.9). The bonding might be direct, as in a C l3 H I fragment, or remote via an electronic framework as in the H-H splittings in ethyl alcohol (CH3CH20H) between the CH 3 protons and the CH2 protons. In solids, the dipolar coupling proves proximity. In liquids, thc dipolar coupling shows up through thc nuclear Overhauscr effect. One thcn distinguishes between bonding and proximity, an important distinction in large biomolecules [7.76] in which a long molecule may fold back on itself. A variety of methods such as spin echo double resonance, spin tickling [7.71], INDOR [7.65J and selective population transfer [7.66J have been employed, to utilize the fact that if I and S are coupled, perturbing the I-spin resonance will produce an effect on the S-spin resonance. All of those methods involve sitting on one line and point-by·point exploring the other lines. leener's discovery of the two-dimensional Fourier transform method converts the approach to a Fourier transform method with all its advantages. In this section we analyze the original Jeener proposal, which now is commonly referred to as the COSY (correlated spectroscopy) method since it reveals which pairs of chemical shifts are correlated by a spin-spin coupling between them. It involves a single nuclear species (e.g. HI) and a pulse sequence
XI(7r/2), XS(7r/2)··· tl ... X / (7r/2), X S (7r/2) ... acquire I(tz) and
S(t2)
,
(7.343) where a single HI of sufficient strength flips bOlh spins. We have already discussed the basic principles of coherence transfer in Sect. 7.25. There we saw that if we are observing the S-spins in time interval two, the density matrix elements (m/-Ie(tl +t2)lmr+) are responsible for the signal. These elements, however, have passed through both (mlmslfl(t)lmrms) and (=fmsle(t>!±ms) during time interval tl. Thus, the I-spin resonance frequencies modulate the S-spin signal during 12 as a function of tt. We can express these facts colloquially by saying Ihe (=f mslu(t)1 ± ms) matrix elemcnts during tl "feed" the (In, ~ Iu(t)!ml ±) elements during t2. Such a description almost suffices 10 describe the pulse sequence (7.343). All one need add is that at t = 0+ both (=fmsle(t)I±ms) and (ml =f lu(t)lml ±) are excited, so that during 12 (ml-lu(t)lml+) is fed by both (=fmslel±ms) and (ml =f lu(t)lml ±). [Note that (mrmslu(t)lmlms) during II also "feeds" (ml -lu(t)lmr+) during 12, as is shown by (7.291) and (7.290), however, since the diagonal elements of e are independent of time they do not by themselves introduce any I I dependence to e.J We could try to carry out an analysis using U similar to our approach to spin coherence. However, Ihis method rapidly becomes quite cumbersome. Indeed, one finds that all 16 elements of U are excited by this pulse scquence! We turn, therefore, to the method of spin operators. 351
Since the two nuclei are identical, we consider a Hamiltonian in the singly rotaling frame, rotating at the angular frequency w of the alternating field. The Hamiltonian includes a chemical shift difference so that il r and ils are not equal. Rather, Ihey are given by (7.344) with ur and us being the two chemical shifts. The Hamiltonian in the rotating frame is still (7.345) Both nuclei experience the same alternating field, H" so Ihat for any pulse the two spins experience the same phase for HI and the same angle of rotation, e. Thus, for example, if there is an Xr(e) there is also an Xs(e). Since we observe both I and S spins, we seek (7.346) with an initial density matrix prior to application of the first pulse given in the high temperature approximation as
(>{O-) =
~(I+ r.~:O(l,+S,»)
(7.347)
Utilizing the time development operators, T, Tr, Ts, Trs, defined in (7.307), we readily get
, H °T, {(l+ + S+)T(t,)XI (,/2)Xs(,/2)T(tl) Z·T
Expressing T(t" as T(t" = Tr(l])Ts(t])T/S(t])
we get T(t,)[yT-1(t]) = Tr(tt)Trs(tt)ciflsS~ll lye-iflsSztl T } (tl )Tr-'(t,) rf l = Tr(t,)Trs(t,)[yTif} (t,)TI- (tl)
T,{ABCj=T'{CABj
r
(7.353)
s
we pull the T '(t2)T '(t2) from the far right hand end of (7.348), giving for the first pan of the trace
Tsl(t2)Tr~ '(t2)([+ + S+)TI (t2)Ts(i2)TIS(t2)..
(7.354)
Then, utilizing eil,S l+e-ilz(};:: l+e i8
(7.355)
and the fact that I and S commute, we get instead of (7.354) (I+e- in ,11 +S+e- insh )T/s(t2)...
(7.356)
so that finally we have
+ S+(t)
J,
s
x X l(rr!2)XS(rr!2)(I: + S:)X '(rr!2)Xi l (rr/2)T-'(tl)
x X,I('/2)Xi 1('/2JT- 1(,,)}
(7.352)
Utilizing the fact that
(I+(t)
(l+(t) + S+(t» = "k
(7.351)
(7.348)
= "'(hHo Tr {1+e- in / 12 + S+e-inS(2) ZkT x [e-ia/,S~h X r(rr!2)X s (rrl2)e ifl ,I,ll e -iaJ,S,ll ly x e+ ia /. S, t1 e -in/ J, 11 Xs(rr!2)X r(rr!2)eiaJ ,S'11]}
(7.357) The expression on the right can be written as the sum of two temls, one containing the I: part of e(O-), the other containing the 5: part. We define these as (7.349) Explicit examination of the two expressions shows that one can be converted to the other if every place one has an [a one replaces it with Sa (a = x, y, z) and every place one has Sa one replaces it with an [a' Thus if one evaluates the teoo involving [: only, one can get the contribution from Sz by interchanging in the first answer Sa's with la's, [a's with Sa's, fir with fl s , and ils with fl[. Therefore, we evaluate just the [z teOll. Then X s(rr!2)Xr (rr!2)[zXi' (rr!2)X = Xs(rr!2)lyX I(rr(2) = ly
S
352
s'
(rr!2)
(7.350)
We have collected inside the square brackets the set of operators which we need to evaluate using the spin-operator foollalism. We saw in our example (Table 7.1) that the operators TJ> acting either on an la or on a product teoo [aSp, at most double the number of teoos, for example transfonning an [y to an ly cos + Ix sin An operalOr Trs acting on a single lerm la or Sp will double the number of tenns, but acting on a product will quadruple the number of teoos. The operators XJ(1r!2) and Xs(rr!2) will change an lz into an [y, etc., bllt not increase the number of teoos. Therefore TIS acting on ly will give two terms, which T J will double to four terms (two single operators, two operator products), which T/S will then convert to at mosl 2 x 2 + 4 x 4 = 12 tenns. (It turns out we get only 9.) They will be of the fonn la, or S{3, or lal{3 after we have reduced any products such as lalp, or [a[pl-y, etc. to the appropriate la" The only temlS which will eventually contribute to (l+(t)+S+(t» f involve either '
e
e.
353
(7.358) These are diagonal in one spin and off-diagonal in the other. All possible spin operators can be expJessed as linear combinalions of the 15 spin functions In. So. IoSIJ (a'" x, y. z; (J = x, y, z) plus Ihe identilY operator. But of these 16, only the eighl operalors (7.359.)
[:nSz. ["Sz. IzS%. IzS"
(7.359b)
will give matrix elements of the fonn (7.358). In fact. using the symbol 0 10 Sland for these eight funclions, (7.357) will consist of a sum of terms Tr {I+O} or Tr {S+O} which vanish if 0 is bilinear. Thus. only the terms (7.359a) of the form [:n. 1y • Sz and
5y
contribute 10 (/+0) + S+(t».
As a result, of the 12 possible terms we will get from [] in (7.357), many will not be of interest. It is therefore useful. as shown by Van den Ven and Hi/bers [7.72] to construct a table in which onc keeps only the spin products, not all the other factors, to find the tenns one needs eventually. then go back 10 get Ihe coefficients. In so doing, we will designale whal we are doing by wriling a sequence of operalors which aCI sequentially in the manner of Sorensen et al. [7.71J. Thus, to indicale that we wish to calculate the effect of lhe operators in the square brackels of (7.357) on an inilial operator I y • we write (7.360) We put these results into a table (Table 7.3) in which over the dividing line belween columns we lisl the operator which has lransfonned one column inlO the next column. If we make our table with 12 lines we should have all the space we need. Comparing Table 7.3 wilh (7.359), we see thai we will gel contributions 10
(£+(1»
Iz
(7.361)
Sz/4
(7.362)
The contribulions 10 I+ can be seen 10 arise from teons which at all stages are either I y or I z • hence diagonal in IllS. Thus they are not fed by tenns which are ever off-diagonal in ms. Therefore, these lenns do nOI involve lransfer of coherence between lhe I-spins and the S-spins. On the other hand. (7.362) shows lhal the tenn Sz/4, which is off-diagonal in ms but diagonal in m/, evolved earlier from the initial 1% to I y , to l z S% to IySy to -1,5". Hence lhe tenn I y • which is off-diagonal in m1 but diagonal in ms. feeds the lenn Sz/4. which is diagonal in m1 but off diagonal in ms. This tenn involves transfer of I-spin 3S4
------- --- -----
I
1.
1.
3
5
-a1,S,I,
X s (,,/2)
-d,S,11
, • , , ""
XI(,,/2)
0,/,11
1.
1.
-I.
I.
,,
I,
I.
I,
I,
3
1,5, _1.5,
1.5,
1,5,
1.5,
[~~:5J x [-~:/.J~
7 8
1.5,
-1,5.
1,5,
10
1,5, +1,5,1,
=
1,5, /.1,5, -(i/2)1,5. -1,5,5, (i/2lf)5. +1,1,5,5. (i/2 11.5,
= = =
5,,/4
polarization to the S-spins. We see Ihat by applying first XS(x{2). then second. X,(7f(1.). we get from I!lS, 10 -I,5,1 via the state IyS". Thi.s operator has malrix elemenls which are off-diagonal in both ml and mS' If mstead we had firsl applied X,(7f{2), Ihen XS(7f(1.). we would have gone from +ly S, to -1,5, to -I,Sll' Thus the intennediate state -I,S, would have had malrix elements which are completely diagonal. This latter Iype of Slate is the one we encountered in our discussion of coherence transfer (7.286). The fact that we go from IlJS, to -I,SlJ no matter which order we apply Xl(7rI2) and XS(7rI2) is an example of our Iheorem of (7.207) that one can interchange the order of the Xl and Xs pulses. We now need to get the rest of the coefficients of I z and Sz on lines 3 and 10 of Table 7.3. We readily find
(r+(,) + s+(t) I, =
-r::;
+ e-insl12 sin (att/2) sin (nll t )2 sin (atV2)Tr{S+Sz/4)). (7.363)
T,{r+r.) =T'{l;J :::I
L
(mlmsl1;lm, 71l S)
=4xi=1
so that
• ,, , "" 5
8
10
Explicit evaluation gives
and 10 (S+(t»I, from line 10
Line
x (e- in/ I , cos (at)(1.) sin (nltl) cos (atV2)Tr{rt l z }
I. from
line 3
Table 7.3. EffecU! of operators from (1.357) on an initial operator I,
(7.364)
(7.365) 355
In writing expressions such as (7.365), Van den Ven and Hi/bers [7.72) introduce a useful nClation which greatly cuts down on the number of lellers one needs to write down. They point out that using this formalism one encounters trigonometric factors such as cos (att!2), sin (atzf2), sin U2[t2), sin (ilst t )
(7.366,)
They write these as
C~
51
== ==
5~ == sin (at2n.) , == sin (nst,)
cos (attn.)
51
sin (nlf2)
(7.366b)
in a notation which is self-evident. One merely has to be careful to remember in the arguments involving a which is not present in those involving l the or ils. Returning to our discussion of (7.365). we now add to this the conuibution from 5~ to e(O-). which we obtain simply by interchanging ill and ils in (7.365). The final result is
"i"
n
(P"(l) + s+(t»
: ~::; {cos (atl/2l cas (at,J2)
(7.367) The various terms in this expression have simple physical meanings. We note fi~t that we have a term involving cos (att/2) cos (atz!2). This term is present even as tt and t2 approach zero. It has the terms
e- inst2 sin (ilstil
and
e-inl12sin (ilstt)+e-insf2sin (illtl)
(7.368a) (7.368b)
which correspond to oscillation near il, during both t I and t2, or oscillation near ils during both tt and t2. Of course the actual eigenfrequencies are ill ± an. and ils±a!2, but if a<: Iill - ilsl the term "near il," or "near ils" is well defined. Near t, :: t2 "" 0, this term is independent of a. It therefore does not require a coupling to exist, and represents an effect of uncoupled spins. as is evident from the fact that the precession frequency is the same (il, or ils) during both time intervals. On the other hand, the second term is proportional to sin (ati/2) sin «(lt2/2). It reaches a maximum when atd2 "" at2/2 "" Tr/2 or
.
(7.370)
The first term corresponds to a spin precessing near ils during the first interval, and near il, during the second. The second term represents precession near ill during the first interval. and near ils during the second. IT one makes a double Fourier transform of (7.366). one gets two types of peaks. From the term involving cos (atl/2) cos (at2/2), one gets peaks near WI"" W2 "" ill or ils. which lie close to the diagonal line WI:: W2 in a (WhW2) plol. From the term involving sin (aid2) sin (at2f2). one gets peaks near the points WI "" ill. W2 "" ils or WI "" nS, W2:: ill. Such peaks immediately show that the spins with chemical shifts il, and ns are coupled. Since the frequencies are always il, ±a/2. not il" and ns±a/2, not ils. the peaks are actually clusters of four peaks. Thus, the diagonal peaks occur at
ill +a/2
W2 "" il, +a/2
wI""il , +a/2
W2""nl-a/2
ill - a/2 w}:fh-a/2
W2 "" nl - a/2 W2""n/+a/2
WI""
x [e-inI12 sin (il/tIl +e-insI2 sin (nstl)] + sin (at In.) sin (atz!2)(e- in,l2 sin (nstj) + e-insl2 sin (n/tl)])
e-inl12 sin (n,tl)
time, t2, after the second pulse that the spins reach the condition of Fig. 7.43c. That coherence transfer is involved is also shown by the fact that this term also involves
WI ""
(7.371)
Likewise, the off-diagonal peaks are actually a cluster of four peaks at (7.3720) and another cluster of four peaks at WI :
fls ± af}.
•
w,:
fI/
± af}.
(7.372b)
If we think of the factor sin (atl/2) sin (at2n.) as modulating signals at il, and ils, we note that the buildup of the cross terms exp (-iil/t2) sin (nst}) or exp(-inst2) sin (il,tl) requires a time ""TrIa. In a typical molecule. a spin I will have coupling constants to both nearby and more distant nuclei. The stronger coupling constants a will produce their cross peaks in a shoner lime than the weaker coupling constants. Thus, by limiting the intervals t} and t2, one can select large or small a's. If, indeed, two spins are not coupled at all, so that a is strictly zero, the cross peaks will never arise.
7.28 Magnetic Resonance Imaging
(7.369) This is the condition we encountered previously in (7.356) as the time which optimizes coherence transfer. This is the time, t" after the first 7(/2 pulse at which the spins have precessed to the condition of Fig. 7.42d. 11 is also the 356
The use of NMR to produce two- and three-dimensional images has, since its invention in the early 1970s. increased at an explosive rate. We refer the reader to the books by MaflSfield [7.77] and by Ernst et al. [7.78] for extended treatments 357
of the many variants. We present a shan treatment of the principles al this point in the text since it enables us to discuss the method of Kumar, Welti, and Emst, which utilizes the two-dimensional Fourier rransfonn approach. Two pioneering papers, published within a few months of one another, are the proposals of lAuterbur [7.79) and of Mall.fjield and Crannell [7.80]. which independently propose use of NMR 10 ronn images. The basic concept is straightforward. We start with a sample which has a narrow NMR line in a uniform slatic magnetic field. To the uniform field we add a second magnetic field which is nonuniform. ideally one with a constant gradient. The field gradient broadens the NMR line. For a constaol gradient (e.g. in the z-direction) the magnetic field component, HI> parallel 10 Ihe uniform field is H~=Ho+z
(OH') fu
(7.373)
so that planes of constant z correspond to planes of constant precession frequency. Thus, a given frequency interval is bounded by two frequencies which correspond to two planes (of constant z). The total NMR intensity in that frequency interval is proportional to the number of nuclei in the sample lying between those planes. Thus the NMR absorption spectrum provides a projection of the sample spin density integrated over planes perpendicular to the gradient direction. From a series of such projections for various gradient directions, one can reconstruct the object. We are considering, then, a uniform field kHo to which there is added a small additional field h(r) which is static in time:
H = kHo + h(r)
(7.3740)
with
h(r) :::: ihz(r) + ihl/(r) + kh~(r)
and
Ihl
(7.374b) (7.374c)
We wish to show first that we can neglect the components h;e and h y perpendicular to kHo. The precession frequency depends on the magnitude of H: w(r) = ~H(r)
.
(7.375)
Since Ho:> Ih(r)l, the effects of ih;e and ih" are merely to rotate H slightly without the first oroer changing its magnitude, whereas h~ changes H(r) to firs! order:
H = V(Ho + h~)2 + =
hi + h~
JH~+2Hoh~+h~+I.;+h~
ll~(r)
= h:(O) + x
==
({)~~: )0 + Y (a~~v )0 + Z(aa~: )0 + ...
h:(O) + xG;e +yGl/+zG;z
(7.378)
which defines the components Gz , G", G: of the gradient G of h~(r) at the origin: (7.379) G = iG z + iG y +kG: The reader may wonder whether we can guarantee that it is possible to generate an 11 with the spatial dependence of (7.378). After all, h must obey the laws of physics (V. It = 0, V x It = 0). We explore these aspects in a homework problem. The answer is that we can achieve (7.378), but for there to be a term such as kxG z for example, there must also be terms in the i and/or i directions. However, since as we have seen the transverse components do not affect IHI to first order, we can neglect them. Let us then define two functions, a frequency distribution function, !(w), which gives the NMR intensity at frequency wand is normalized to satisfy
J
f(w)dw = 1
(7.380)
and a spin density function e(r), which gives the number of nuclear spins in a unit volume at point r, and which is likewise normalized so that
J
(7.381)
(,(r)d'r = 1
If, then, the gradient G is oriented in the i-direction,
G=
(7.382)
Z'G .
planes i = constant are planes of constant precession frequency. Since the magnetic field changes by G dz' between the planes al i = constanl and (z' + d.z') = constant, the precession frequency change, dw, is (7.383)
dw=;Gdz' Therefore
!(w)dw = dz'
f'lJ
dx 'd y"",x,y,z I"J I I I)
(7.384)
%':::consl
(7.376)
Neglecting the terms quadratic in the components of h, we get
H = HoJI + 2h~/Ho ~ HoO + h:IHo) = Ho + h~
What spatial forms can h: take? If we simply consider it to be a gen~ral function of r, we can expand it in a power series in z, y, z about a conventent origin. We assume we have shaped the field so that only the lowest terms are needed. Then
(7.377)
The integral in (7.384) is the projection of the spin density funclion e(x', yl, z') on the z'-axis. Thus, utilizing (7.383), "(Gj(w)::::
11
dx1dy'e(:i,y',/)
(7.385)
~/=consl
358
359
Since this projection depends on the direction of G, and since the magnitude of f(w) depends on the magnitude of G, we put a subscript on f(w) :
.."Gfc(w)
=
JJ
dx'dy'e(z', y', z')
(7.386)
:r'=const Since fc(w) can be measured, we can view the integral on the right as known experimentally as a function of z'. We can think of making a variety of measurements 10 delennine the w dependence of fc(w) for a variety of different G directions, thereby getting the projecled spin density for any desired projection axis, z'. It is now easy 10 show that such measurements enable one to reconSmJct e(z, y, z). Firsl we take the experimental dara for a given z' and fonn the Fourier transform of the data:
ek =
J
e-ik:r'dz'
JJ
dx'dy'e(z',y',/)
problem then becomes one of finding a suitable numerical approach for getting the most accurate value of the integral from the data as given. Of course, in the process of selecting the approach, one gets guidance as to whal data to collect. The approach of reconstructing e(r) by Fourier transforming the spin density distributions is one approach to finding e(r) from data for fc(w). Lawerbur [7.79] demonstrated a simple graphical means, and pointed out that Ihe solution to the problem of getting e(r) from spin-density projections was well known from Olher areas of science. Another experimental approach was proposed by Mansfield, based on 0bserving free induction decays. If one applies an X(ll'"{l) pulse to the spins and observes the free induction decay in Ihe reference frame rotating at wo, the complex magnelization is
(M+(tl) = (M.(t)} +i(M,(tl)
(7.387a)
= iMo
where = iMo
k = H'
(7.387b)
Utilizing (7.373) wilh Wo
ek = exp(ikwo!"'(G>
== "'(Ho
exp(-ikw!"'(G)fc(w)dw
.
(7.388)
If we know ek for all k, we can get e(1") by the relalionship 1 l'\r = (2ll'")3 -J )
j eik·.ek. cr"k· .
(7.389)
'. k,
.'11:.7.46. The paths in k-space (k = ki') eXI>lored by applying the gradient C = G i , in a variety or directions i' in a rhUle. The various directions 91, 92, ... ,9,. correspond to different directions i;,
-,
" 360
(M+(t» = iMo
so that (M+(t» =iMo
In practice, since each direction k is explored in a separate experiment, ek. is never known experimentally for k as a continuous variable. For example, if one has a two-dimensional object, the various k's would be chosen to lie in the plane of the object. Then the k values explored could be represented as in Fig.7.46. It is seen that they represent radial lines out to some maximum k, and k max • Since performing an integral such as (7.389) is a numerical operation, the
9,
e(z', y', z')e-iw(z' ,JI,z')ldz'dy' dz' f!{z', y', i)e-i.,.G:r"d:i dy' dz'
(7.39Oa)
which can be rewritten to make ils meaning clearer as
and (7.386) we can express (7.387) as
J
J J
i:;,
J J
JJ JJ
dz/e-i.,.G:r'1
dz'e- iq ''''
dz' dy' f!.z', y'. z')
dz'dy'(!{Z',y',z')
(7.3901»
(7.391)
where
(7.392)
We can rewrite \1.391) to emphasize that on the left we are observing
(M+(t» for a particular gradient direction i', and rewrite the right hand side to bring our its meaning to get
(M.+(t»j;' = je-iq • .,. e(r)d3r
.M,
(7.393)
The expression on the right of (7.393) is identical to that on the right hand side of (7.387a) if one replaces q by k. We will henceforth replace q by k. Thus, in the presence of a static field gradient, the NMR signal following a 7r{l pulse sweeps out in time the Fourier transform (?k' where the direction of k is given by the direction of the field gradient, and each point in time corresponds to a magnitude of wave vector given by (7.392). Clearly, the free induction decay is also producing f:!k along radial lines in k-space as in Fig. 7.46. In his initial paper, Mansfield [7.80] proposed using the free induction decay, giving (7.393). Since his initial idea related to crystallography, he explored the case that the spin density is given by a periodic lauice, and demonstrated the imaging scheme by making a macroscopic layered sample to simulate atomic periodicity.
36.
In Fig.7.46 we display a set of k trajectories for a planar sample. Even for nonplanar samples, it is convenient to collect data from planar slices in me sample, a sequence of slices thus providing the full three-dimensional object The methods of selecting well-defined slices are pan of the present an of imaging. We refer the reader to (7.77] for example. A simple approach to generate a slice perpendicular to the z-direction is to start by applying a gradient in the z-direction. If one applies a pulse at a panicular frequency WI, the spins at coordinate z such that
which is the Fourier transform in the plane z of the spin density. Since f z is fixed while recording (M+(t», the data are for a fixed kz • but for a continuous k y . They may be represented as in Fig. 7.47. They thus sample (}k for the z-slice along parallel lines in (k;r:. kll) space. In a full three-dimensional version, one applies the initial rl2 pulse in the absence of gradients. Then. one applies
(7.394) are perfectly at resonance. An X , (7r!2) pulse will then rotate them into the x-y plane. Spins within a Ll.z given by
(7.396,)
G=iG;r:
fo<
I.
G=iC" G= kG r
for
Iv
for
I.
(7.400)
recording the free induction decay during t r • Then
(M+(t» = iMo
(7.395) will also be flipped. but spins farther away in z wiJl be less effectively tilted into the :r-y plane. By in fact shaping the time dependence of (HI) I one can improve on the z-selection. In any event. if one now turns off Gr. one has produced an initial transverse magnetization for all x. y within Ll.z. One can then apply gradients in the :r-y plane to record the free induction decay for any direction k = G in the x-y plane. All planar imaging methods must initially carry out some such precedure to define Ihe slice Lb. Kwnar et al. {7.81] proposed a two-dimensional (or three-dimensional) Fourier transform approach to collecting data which has been adopted by many workers in the field. We start by describing the two-dimensional version in which one first exciles a slice in the z·direction. One then applies a gradient
G=iG z
J
e-izGzlze-iIlG,I'c-irG.I. p{x. y. z)dx dy dz
(7.401a)
with (7.4Olb)
""d t z = t - (t z
+ ty)
(1.40lc)
This approach gives one the Fourier transfonn of u(r) lines parallel to the k r axes fonned by the intersection of the planes k>; = constant and kll = constant (Fig. 7.48). If one is interested in knowing the density pattern in a slice of known %. it is much faster to collect the data using a z·gradient to define the slice excitation and then work with two-dimensional transfonns. An exceedingly rapid way to sample all the points in one slice within a single sweep was proposed by Mansfield (7.82J and called by him the planar echo method. In the planar fonn of this method. one first selects a slice (say
for a time called f;r:. then a gradient
"
(7.396b)
G = iCy
for a time called t y • During t y • one records the free induction decay. Using (7.170) and (7.390) one has, neglecting relaxation. in Ihe rotating frame
(M+(t» = iMoL\z
J
e-i;r:,..Gzlze-iyG,f, e{x, y, z)dx dy.
FIg. 7.47
"
Fig. 7.48
(7.397)
where t = t y + t;r: and where M o is the tolal magnetization of the sample. Defining (7.398)
k;r: = '"'{G;r:t;r:
= L\z
J
e-ikz;r:e-ik,1I e(x, y, z)dx dy
'.
F1a. 7.47. The palhs in k-llpace explored by the twu-dimerl!lional meLhod or KW1«Ir, Welli, andErltSl. For a particular lz, k z is fixed. As the Lime f, grows (7.397), k, grows as shown
one can write
(~;:t»
"
(7.399)
Fig. 7.4&. The path explored in k-space by the and Ernst
th~imensional method
of KIU1tDT, Wtlli,
363
t
at fixed z), then applies a weak steady gradient in the z-direction, and a strong gradient in ~e y-direction which is switched ahemately between two values +GO and -Go (FIg. 7.49). Thus, following slice selection one has G =
iG z + JGO/CO where
j(t)=+l
O
-I
for
Tf,
+1
for
3T,,
(7.403)
and so on (Fig. 7.49).
,.,---==='.
In order 10 ~asp the effect of such a time-dependent gradient, it is useful to follow our earlier procedure of defining a wave vector k. The approach we then take was developed by King and Moran [7.83J, BrOWII et aI. [7.84], Twieg [7.85), and Ljunggren [7.86). Since the precession frequency w and the angle a spin has precessed ,p, are related by
d4> d'
w=-
(7.404)
when the gradient is time dependent, we have t
J [-i J
W(t')dt'] eCr)d3r
(7.405.)
o = iMO
Therefore, defining
364
J
ex p [ -i'Y""
J
G(tl)dt'
t
J
G(t')dt']u
o
(7.406)
o (7.407) Thus. as t develops, k(t) moves Ihrough k space on (he trajectory given by (7.406). (.M+(t»/iMo gives f!k(t) for the successive values of k through which (7.406) takes the system. Examining the gradient of (7.403) or in Fig. 7.49, we see that the weak COIlstant component Gz gives a component kz which increnscs slowly proportional to lime. The strong gmdient G y builds up a large phase angle z,Gzt from 0 to Tb' but the sign reversal of G z which then occurs leads to this phase accumulation unwinding during the next interval 10. Thus at t = 2r" the various planes :t = conSlant gel back in phase, producing an echo that is (7.408)
fo'ig.7.49. Thc limc dependcncc or Ule lhree ~rlldicllls G z , G" aud G. cmployed III lhc planar echo Illelhod of Mmufrdd
exp
== ,
(7.402)
for
(M+(t» = iMo
k(t)
(7.405b)
In k-space, the vectorik",(t)+jl.-J/(t) follows the Inljectory shown in Fig. 7.50, in which we have assumed both G", and Go are positive. As can be seen, this method explores positive and negative k y for JX>sitive kz . One can get the resuhs for negative kz by several methods. One is to apply a strong negative G z for a short imerval after slice selection, and before applying G y • This process would move k to a point on the negative },:z-nxis. Then one begins the sequence of (7.374n). Another approach would be to utilize the mathematical result that
(M+( -k» iMo
=
(M+(k»)' iMo
(7.409)
which can be scen to be true from (7.407). This is the statement that with quadrature detection the signal of one phase gives the cosine transform, the other gives the sine transform. The practical result is that one need only collect quadrature data in one half-plane of (k"" ky). This remark of course holds true for the schemes we described earlier. We see, therefore, that the "planar echo" method samples all the half-plane in each cycle, and thus contains the infonnation for total spin-density reconstruction in each sweep. A more "physical" description of the planar echo method provided Mansfield's original motivation. He thoughl in terms of a discrele laltice (as of atoms in a regular array). Suppose we imagine an apple orchard with trees planled in equally spaced rows and columns. SupJX>se some of the trees have died and been cut down. If one stands at the end of one row and looks down that row, one cannot tell whether any trees might be missing from it. However, since one sees the next row at an angle, one can readily spot any vacant site. This phenomenon is illustrated in Fig. 7.51 from Mansfield's book [7.77]. To utilize this principle in imaging, one must produce a discrete lattice, and view it at an angle. TIle imag· 365
8. Advanced Concepts in Pulsed Magnetic Resonance
k,
0
k.
0
0
0
0
I
0
Gy
0
Hahll's discovery of spin echoes [8.1) demonstrated thaI one could defeat the effect of magnet inhomogeneity in obscuring the true width of magnetic resonance lines bec;lUse the echo amplitude M(2T) decayed exponentially willi 2r!T2. where T is the time between pulses. However, if the magnet inhomogeneity is large, and if the diffusion rale of the nuclei sufficiently great,
0 .y(j,t~
M(2r) = M(O) "p(-2r/T,)exp [-
I Fig.7.S0. The {mth in k-spa.<:e explored in the planar echo method of M(lflSfitld
Fig. 7.51. Projections of a discrete matrix filled to form the letter P. ProjectiolU in G z or 0, alone do not contain sufficient inforllllllion to uniquely determine the original object. Projection orthog<mal to the special gradient direction G. conlains all information necessary to complelely'del.crmine the original objed distribution
ing scheme of Fig. 7.50 has parallel scanning lraces which corresponds to a spatial wavelength
2.
, __ 2r __ k",
~1'GJl70
apan in k-space,
(7.410)
A
21Gyf],
This wavelength imposes discreteness (i.e. "rows'') on the spin density function
elr). Another way of saying this is that the echo train produced by the oscillatory gradient G y has a period 2Tb. hence corresponds to observations al a spectrum of frequencies spaced apan by angular frequency Llw of
L\w
"fGy
Dr']
(8.1)
where D is the diffusion constant and DHIDz the gradient in static field. In some instances, then, the r 3 leon might obscure the T2- Or, conversely, if DHlfJz is known, the T 3 term gives one a method of measuring D. The derivation of the diffusion tenn is given in Appendix G. Carr and Purcell [8.2] invented a clever scheme for eliminating the T 3 tenn when desired. 'Their proposal initated a class of experiments in which a sequence of pulses is applied 10 the resonance. In this chapter we discuss first the Carr· Purcell concept, then an ingenious modification invented by Meiboom [8.3]. and finally tum to some of the most intricate pulse melhods yet invented which have the remarkable propeny of enabling one to eliminate the dipolar broadening of resonance lines in solids. 11lese concepts were introduced by Waugh [8.4J and Mansfield [8.5], and have been extended by Vaughan [8.6J. as well as a group of very talented resonators who have spread Qut from these initiators.
(7.411)
Bm L1w implies a ..1z of LIz = -
(1 ~:)' ~
8.2 The Carr-Purcell Sequence
1 Llw = 2r,
= -
I
-
1
(7.412)
""IC" 27"b
Each frequency corresponds 10 a row of trees. The weak z-gradient produces the "view al an angle" shown in Fig. 7.51. 366
8.1 Introduction
We begin a discussion of the Carr-Purcell sequence by llssuming we call neglect the T 3 diffusion tenn. To measure a T2 by the conventional spin echo, one must make a sequence of measurements, each al a different value of T. The envelope of the echo amplitude as a funClion Of2T gives one T2. Carr and Purcell pointed out that the entire envelope could be obtained at a shot if one applied the proper sequence of pulses.
367
Suppose at t::: 0 one applies a 1r/2 pulse in which HI lies along the +x-axis in the rotating reference frame with the rf frequency w exactly at the Larmor frequency woo Such a pulse turns the magnetization Mo to lie along Ihe +y-axis. If one applies a 1r pulse at t ::: r, also with HI along the +x-axis, an echo is fanned at 2r with the magnetization along the -y-axis. If now one applies a 1r pulse at 3r, another echo will foml al 47, this time along the +-axis. In this manner, successive 1r pulses at (2n+l)r(n::: 0, 1,2, ... ) foml echoes at (2n+2)r, the echoes fanning along the -y-axis for odd 11, and the +y-axis for even n. Since all components of the magnetiz..1lion in the x-y plane are decaying exponentially with time constant T2, the echo sequence likewise decays exponentially with T2. A sequence of such pulses and echoes is shown in Fig. 8.1, with Ihe sign, positive or negative, showing whether the echo fonns along the positive or negative y~axis. The behavior we describe enables one to obtain the entire echo envelope in a single pulse train, clearly a convenience. If one is studying a weak signal, so that the signal-to-noise ratio is imponant, the Carr-Purcell sequence is a tremendous advantage as we explain. The noise depends on the bandwidth of the apparatus. One limits the bandwidth so as to pass the echo signal without undue attenuation. Clearly the same bandwidth can be used for either a Carr-Purcell train or a convelllional echo. After each single echo, one must wait several times TI for the system to recover before measuring the next echo. However, in the time il lakes to get one echo, we can obtain the full envelope with a Carr-Purcell sequence. If we take N echo signals to define the echo envelope, we could obtain N Carr-Purcell pulse trains, so each Carr-Purcell echo is recorded N times. If the spin echoes are taken to correspond to the times of the Carr-Purcell echoes, the Carr-Purcell train will have a better signal-to-noise by Jiii. We have remarked that the Carr-Purcell pulse sequence reduces the effect of diffusion on the echo decay. The reason can be seen as follows. If there is no diffusion, a spin dephases during the interval r following each echo and rephases
,
o
.
2,
_---- . -------
------ ------
),
M(2r) "" Mo exp [ _ D
------
---------- ---------
6,
7,
--------
"
(7 ~~)
2
~r3] exp(-2rIT2)
== Moa
(8.2)
When the echo is formed, we have recreated at 2r the situation which existed at t ::: 0, except the magnetization is reduced by the factor a, and points along the -y-axis instead of the +y-axis. Therefore we can consider the magnetization at 27 as another initial condition for applying the argument of Appendix G for the next interval from 27 to 4r with a pulse applied at 3r. The result is (8.3) M(4r) ::: M(2r)a::: Moa 2
If we have n cycles 2r, we therefore get (8.4)
M(n2r) ::: Moa" Therefore M(n2r) "" Mo exp [ -
(Y ~~y D(n27)~72] exp( -n2r1T2)
(8.5)
Thus, if we change r but hold n2r constant so that we are comparing two different echoes of the Carr-Purcell train a fixed time apart, we can vary the faclor involving diffusion but we do not vary the faclor containing T2. Hence we can make the effeci of the diffusion tenn negligible. It is clear thai what is involved are two facls: 1.
5,
2.
each echo cycle of the Carr-Purcell train reduces the magnetization by a factor 0', the relative contribution of the diffusion tenn compared to the T2 term depends on 7 .
Echo clll'dope
8.3 The Phase Alternation and Meiboom-Gill Methods
----------
Echo CIII'i'lop"
Fig.S.1. The Carr-Purcell sequence and tlte resulting echoes. As des<:ribed in the text, a positive echo forms along the +y-axis in the rotating frame, a negative echo forms on the -y-axis
368
during the interval after Ihe 1r pulse. It is the fact Ihal diffusion makes rephasing imperfect which causes the next echo to be smaller. In Appendix G we show that if a 1r/2 pulse at t "" 0 tilts the magnetization Mo into the x - y plane, a 1r pulse r later produces an echo at 27, M(2r), smaller than Mo owing to diffusion by
A Carr-Purcell pulse sequence may contain many pulses. If the pulse rotations deviate slighliy from 'IT, there are cumulative effecls which may become large. We illustrate in Fig. 8.2. (This figure is drawn for a negative "'/' hence involves right hand rotations.) For simplicity we take the initial pulse to be a perfect 1r(l pulse about the +x-axis, and we consider diffusion effects 10 be negliglible. At 369
J ,
(,)
-/2 M.
-----'7---r x
(b)
---0- -----
f--_ ,
'......
.• -------
r" O'
Meiboom and Gill [8.3] were the first people to find a solution to this problem. They likewise used a phase coherent method, but introduced a 90" phase shift into the radio frequency field between the initial pulse and the subsequent 'If pulses. Thus if HI for the 1f/2 pulse were along the +y-axis, Hl for the :IT pulses would be applied along the +x-axis. As a result, the echoes are all formed along the +x-axis for negative 'Y' We leave it to the reader to show that if the pulse is 1f - 5 instead of 11", the error does not accumulate.
7rn.
"\
y
"'./
, (d)
8.4 Refocusing Dipolar Coupling ---.--'-"'-"'-,,---- y
''''(31)'
Fig.&.2a-e. The effect of ,an imperfect If pulse. The initial ../2 pul$e, assunl«! perfect, ro~aleti Mo to the - y-axIS at I = 0+. At t = T - • just bo;orore the .$eOOnd pulse, a typical spm h~ Pl'('(;~ an .~~Ilt 8 in the z - II plane a.....y from the -v-
makes an angle 26,p wIth the z -" plane. In Fig. 8.2e we see that if the III of the second pulse lies along the -x-axis, the rotation direc:tiOll is reversed, so the _(,.. + 6.,> rOlalion restores the spin to ils orientation at t ::: 0+
t ::: 7 - the spins will have fanned oul in the x - y plane. a typical spin making an angle 8 with the -y-axis (for negative ")'). At t "" T we apply a ow + 6¢ pulse about the +x-axis where 6¢ gives the deviation from being a perfect 'If pulse. For a spin having 8 "" the spin lies along the x-axis. It is not affected by the pulse at t"" T. A spin for which 8"" 0 would rotate as shown in Fig. 8.2c. It will lie an angle 6¢ above the +y-axis in the y - z plane where it will sit until 2T later when the next + b¢ pulse is applied. The result is shown in Fig.8.2d. The spin now makes an angle U¢ with the -y-axis. Each successive imperfect 1'1" pulse adds another 6¢ to the deviation. The cllmulative effect becomes exceedingly serious in praclice because the HI is never uniform throllgh tile sample. COllsequclIlfy thOllgh part o/the sample may have a ow pulse, other part.r do f1ot. These sons of troubles have major implications for all multipulse sequences ~uch as those described in the later portions of this chapter. One simple approach IS to apply the H J for the ow pulses alternating along the +x- and -x-axis. Thus the sense of rotation about the x-axis reverses in alternate pulses. The result is shown in Fig. 8.2e. Instead of getting a cumulative rotation eITor of U¢ as in Fig.8.2d, the cumulative error is zero!
7rn..
7r
370
11le dipole-dipole coupling between neighbouring nuclei provides valuable in-
formation for many purposes, but on occasion it is a problem. For example, it may obscure chemical shifts, and it may cause free induction decays to be soonlived and thus difficult to see. We tum now to several interesting approaches to eliminating or effectively reducing dipolar coupling. For a group of spins an ordinary spin echo [e.g. an X('IT/2) ... T ... X(:IT) sequence] refocuses the dephasing which arises because they are placed in an inhomogeneous magnetic field. Since the magnelic dipolar coupling among neighboring spins is in some way analogous to a magnetic field inhomogeneity, one might ex:pect that the same spin echo pulse sequence would also refocus the dephasing resulting from magnetic dipolar coupling. If the dipolar coupling is between different nuclear species (e.g. HI and C I3 ). such an echo sequence does refocus the dipolar coupling. If, however, the coupling is between like nuclei (e.g. HI with HI), the echo does not work. It is easy to see why. The 'IT pulse inverts all the neighbors, so that a nucleus which was precessing more rapidly than average in the first time interval T, finds its neighbors inverted, and thus precesses more slowly in the second interval. The usual echo works because the 11" pulse converts a precession phase lead into a precession phase lag, of equal size. If at the same time the rate of precession changes, the spins do not come back into phase. Nevertheless, the fact that a well-defined relationship exists between the precession frequency when a neighbor points up versus down makes one feel that there should be some way to undo the dephasing which the dipolar coupling produces. In the following sections, we explore these ideas.
8.5 Solid Echoes Powles and MallSfield [8.7] discovered that the free induction decay of a coupled pulses pair of identical spin ~ nuclei can be refocused by applying a pair of shifted in phase by ow/2 with respect to each other. Such a sequence might be denoted by
7rn.
371
X('rr{2) .. . r . .. Y('II"/2) . .. tt
(8.6)
with the echo occurring when tl := , . This sequence has come to be known as the solid echo. Powles and Mansfield demonstrated it experimentally for CaSO'1 ·2H20, and showed that it followed theoretically. The pulse sequence perfectly refocuses spin nuclei which interact only in pairs, but does not refocus larger groupings of spins perfectly. In this section, we explain the echo for coupled pairs. This same pulse sequence refocuses the first order quadrupole coupling of a spin I nucleus for reasons explained in Appendix H. It is therefore imponam for deuterium and N t4 NMR. The existence of echoes related to quadrupole splittings was first demollStrated by Solomon [8.8] for a case of I := ~. It appears to have been Davis et al. [8.9] who recognized that for the 1:= I case sequence (8.6) gave perfect refocusing. The fact that one needs exactly sequence (8.6) may seem surprising at first glance. Why does one need to phase shift the second pulse by 11:/2? Why does one use a '11"/2 pulse instead of a 'II" pulse? To help give a feel for the answers to these questions, we shall derive the result Ihree ways. The first method is a semigraphical one, utilizing a special value of r. The second derivation generalizes the first for arbitrary r. The third derivation utilizes the spin operator method. We start with the simple physical picture. However, we must present it rather carefully. Consider a sample consisting of pairs of nuclei. That is. each nucleus has one and only one neighbor which is sufficiently close thai only the dipolar coupling to it matters. In the presence of a static field, the nuclei are weakly polarized, there being a slight excess with magnetic moments parallel to the static field, H o. If there are N nuclei and if we denote by p+ and p_ the probabilities for a given nuclear moment to point either parallel (p+) or antiparallel (p_) to H o, we have then that there are N+ and N_ nuclei pointing parallel or antiparallel, respectively, given by
4
(8.7) The population difference, n, then obeys n:=N(p+-p_)
(8.8)
p+ and p_ are related by the Boltzmann factor p+/p_
:=
e"tldfo/kT
(8.9)
so that, if ",(hHo «kT, P+:= ~(1 + "'(fiHo/2kT)
p_
:=
~(1 - "'(hHo/2kT)
(8.10)
Note that the expressions for p+ or p_ do fU)( depend on the orientation of the other nucleus in the pair, because we have assumed that the only field acting on a nucleus is Ho. This assumption should be valid if Ho d> "'(1lIr 3 , where r is the distance between neighbors. Now, it is only the excess of nuclei N+ over those N_ which can give rise to a NMR signal. To follow what happens, we need 372
only follow what happens to the excess, since Ihe remaining nuclei carry no nel polarization. Let us therefore in a thought experiment reach in at random to examine the pairs. We take n/2 pairs in which both nuclei point up, and n{2 pairs in which one nucleus points up, and the other points down. This will give us a net n nuclei pointing up. The olher N ~ 2n nuclei in the sample must then be equally pointing up and down, hence be unpolarized. We neglect them. Now, among the pairs with one nucleus pointing up, Ihe other pointing down, we label the up nucleus blue, the down nucleus red. Among the pairs with both nuclei up, we also label one blue, the other red (it does not matter which of the pair we choose for which color). We thus have n{2 red nuclei with spin up, and n/2 red nuclei with spin down. Then the blue nuclei give the magnetization, since the red nuclei have no net magnetization. We represent the spin-spin coupling by a term 1t'spin-spin of the form (8.11)
!
For pairs of identical particles of spin this is exactly equivalent to dipolar coupling, as explained in Appendix H. We assume that apart from 1t'spin-spin we are exactly at resonance for any pulses we apply. Then, following an X(1r{2) pulse, the blue spins lie along the +y-axis in the rotating frame. Now, although we have flipped both red and blue spins by 1r{2, the red spins at this point have no net polarization and thus their density matrix remains diagonal with the two diagonal elements equal to each other. Thus, we can describe the situation at t := 0+, just after the pulse, as in Fig. 8.3b, in which we represenl the blue spins by vectors I and 3, lyir]g along the y·axis, and the red spins by vectors 2 and 4, one pointing up, the other down, along the z-axis. Spins I and 3 will now precess at rates -a!2 and +a/2 (in the left-handed sense) respectively about the +z-axis. The explanation of how the echo fonns becomes very simple for a particular value of " the value which makes arl2 := 1r12 (Le. a, := 1r). For that time, the spins will be aligned as in Fig.8.3c. At this time, the y('II"!2) pulse will produce the situation of Fig. 8.3d. Now, during the next time interval" spins 1 and 3 will not precess (they lie along the z-axis), but spins 2 and 4 will precess at rates +a/2 and -a/2 respectively so that at t:= 2" 2 and 4 will lie along the +y-axis. An echo therefore occurs at t "" 2,. It is interesting to note that in this process we have transferred the net polarization which was initially in spins 1 and 3 10 spins 2 and 4. However, we have not made an unpolarized system become polarized. Thus, we can remain confident that the other N - 2n spins, which we have said are initially unpolarized, do nOI become polarized by the echo sequence. If one chooses a , other than Ihe one for which arl2 := '11"12, the explanation becomes more complicated. Then one needs to utilize the ideas of Sect. 7.24 in which we discuss the time development of one spin when it is coupled to another spin which is in a mixture of the spin-up and spin-down eigenstates. By choosing a,/2 := 'II"!2 we have made this mixture be in fact pure eigenstates. 373
However, a true echo means thai the refocusing condition is independent of the strength of the source of the dephasing. Thus. if spins are dephased by an inhomogeneous magnet. all the spins get back in phase at the echo. independent of how far the inhomogeneity has shifled their resonance from the average field_ In our case. that means the existence of an echo should be independent of a, hence of the PrOOuC[ aT, hence of T. Let us therefore reexamine the situation for a more general T, using Fig. 8.4. We start (Fig. 8.4a) with the magnetization along the z-axis, and apply an X('K(l) pulse to prOOuce the situation in Fig. 8.4b. A time T later the vectors I and 3 have rotated in the % - Y plane through angles (aT(l) and (-aT(l) to produce Fig.8Ac. Then the Y(1I'/2) pulse produces the situation of Fig.8Ad in which magnetization vectors 1 and 3 are in the y - z plane, making angles -aT(l and arl2 respectively with respect to the y-axis. In Fig. 8.4e we show the projections of the four vectors onto the x - y plane. Spins 2 and 4 lie entirely in the plane. We define their lengths as Mo. Spins I and 3 have projections in the x - y plane of Mocos(aT(2.). Now, using the concepts of Sect. 7.22, we realize that each of the four magnetization vectors is coupled to a spin which is in a mixture of up and down states. Since the quantum states of spins 2 and 4 are equal mixtures of up and down. spins 1 and 3 will each break into two counter-rotating components of equal amplitude (Fig. 8At). Spins I and 3 have z-components of -Mo sin(ar(2.) and +Mo sin(aT(2.) respectively. Thus they have an excess of m = (for spin 1) and m = +i (for spin 3). We need to calculate what fraction of their stales correspond to spin up and spin down. Denoting Ihe occupation probabilily of slate m for spin i as (mIUilm) we have that
h: O'
, y t: 0'
l 4
,
,IT: "
,
Idl 2
-!
y I: '['
,
,
(8.12)
from nonnalization. The z-component of magnetization of spin I,
y
lei
2
,
t :
M" = Mo (~I",I~)
2'[
expressing the fact Ihat if the spin is entirely in respectively. But
I'Ig.S.Ja·e. The fonnation of a solid echo for two pairs of spin!! (I and 2, 3 and 4) for a particular value of the time T" at which the refocusing pulse is applied. (8) The thermal equilibrium magnetization at 1 0- showing the vector sum of the four magnetization veclors. Spins 1 and 3 are the blue sl)ina, spins 2 and 4 are the red spins. (b) Just after the X(Tr/2) pulse (t 0+), the blue spina (1 and 3) lie along the y-axis while the red spins (2 and 4) can be taken to point along t.he positive Il.Ild negative ",-axC!! respeclively. (c) The magnetization vectOrs at times L = T- just before the Y(lr/2) pulse, for the particular T such that (JT/2 lr/2 (or or = lr). (d) The magneti:u.tion vectors at time 1 1+ immediately llfter the Y(lr/2) pulse. (e) The magnetization vectors at time t 2r, showing that there is a net magnetization IIlong the y-axi!! from spin!! 2 llnd 4. Note that initially spill!! I and 3 carried the mllgnetization. At t = 2,., the magnetiZll.tion arises from spins 2 and 4
=
=
=
- Mo (- ~Ied -~)
=
=
M~l
= -Mosin(ar(l)
M~t>
is (8.13)
i or -i. M
z1
is Mo or -Mo (8.14a)
so we get
(~Ied~)
- (-~Ied-~)
=-s;n(aT/2)
(8.I'b)
Solving (8.12) and (8.13) we get Well~) = ~[I - s;n(aT!2)]
(- ~Ied -~)
=
1(1 + s;n(aT!2J]
(8.15)
In a similar way, we get 374
375
,
Fig.8.4 Caption see opposite page
(- ~I",I-~)
lal
, ,
Ibl
,
,
,
2
2 all2
--
,
,
~o [1
,
4
1
11
/
/
,
lei
,
(8.18)
N~O [1
- sin(ar/2)] sin(ar/2)
~O [sin2(or/2) + sin2(ur/2)]
(8.19)
Therefore
L
t~,'
No
4
M y (t+2) = Mo[sin 2(or/2) + cos 2(or/2)] = Mo
,
Igi
(8.20b)
,
M
(os(aTl2)
•
Fig.8.4a·g. Formation of a solid ocho for two pairs of spins (I and 2, 3 and 'i) for a general value of the spin-spin coupling conslant u. (a) The thermal equilibrium magnetization at t = 0- showing the vector sum of the four magnetization vectors. Spins 1 and 3 are the blue SpillS, spills 2 and 'I are the red spins. (b) The magnetization vectors at 1=0+ immediately following the X(1r/2) pul$(!. (e) The magnetization vedors at time t = r-, just before the Y(1r/2) pulse. (d) The magnetization vectors at time t r+, immediately after the Y(tr/2) pulse. (e) The projections of the magnetization vectors on the :l' - Y plane at time t = r+, immediately after the Y(tr/2) pulse. (f) The projection of the magnetization components of a spin 1 on the 2: - Y plane at later times, showing how its Illagnetization breaks into two counter-rotating components whose amplitude is determined by the orientation of spin 2. (g) The projection of the magnetization components of a spin 2 on the z - !I plane at later times, showing how its magnetization vector breah into two counter-rotating components whose amplitude is determined by the z-components of spin 1
=
spin 2
spin
("'T---'
(8.20')
By a similar argument
Ho
-f
~o [1 + sin(or/2)] sin(or/2)
=
=
_-'-;:,-...L__ ,
If I
M y2
_
~--Mo(os(aTl2l--- 3
1
~o cos2(aT/2) + ~o cos2(aT/2)
,
,lTn /
M y1 =
,
,
Idl
- sin(ar!2)]
as shown in Fig. 8.4f. We now add up the y-components, My; (i = 1,2,3,4) of the spins at time -r after the Y(7l"/2) pulse:
"-k'::..o" )3 aT 12
,
(8.17)
while the counterclockwise component of 2 will have amplitude
,
4
,
I,}
~o (1 + si,(arl2)]
3
,
(8.16)
= HI - si,(arl2)]
Thus', as spin 2 precesses, it will break into two counter-rotating components whose amplilUdes are detennined by the relative amounts of spin 1 in the up and down states. That is, the component of 2 which rotates clockwise (looking down on the x - y plane) which comes from the down component of 1 will have amplitude
,
376
W",I~) = ~(I + si,(arl2»)
M. T
l1-sin(aT1211
377
This result is independent of a, therefore corresponds 10 a real echo. Indeed, (8.20) shows that in a powder sample (in which a would take on different values for each crystal orientation if it were representing magnelic dipolar coupling), spins of all the crystallites would refocus at time 2r. This exercise has been long, laborious and exhausling, but it does serve to show how coupled precessing veclors behave in quantum mechanics. Now we lum to a much more elegant approach to analyze the same problem: use of the product operator method. It has the advantage of giving us a simple way of figuring OUI what the second pulse should be in order to produce the echo. (See Seci. 7.26. See also [8.10, lIn. We start with a system in thermal equilibrium and work in the rotating reference frame. The density matrix JUSt before the first pulse (t = 0-) is (8.21)
where A is a constant. Following the pulse X(1r(2), the density matrix becomes
~~=~+~
~~
The density matrix at time t = r- JUSt before the Y(ll"/2) pulse is then e(r-) = e- ia1• S ... e(O+)e+iol.S.T
Now, utilizing (8.23) we have e+iolzS... e(O+)e-iaI,S, .. . = (Iy + Sy)cos(ar/2) + ([ZSl + I z S;l:)2sin(ar/2)
(8.29)
which, using (8.24), is equal to
Re(r-)R- 1
(8.30)
and using (8.23) gives
R[(ly + Sy}cos(ar/2) - (l;l: + S;l:)2 sin(ar/2)]R- 1
(8.31)
Comparing (8.29) with (8.31) we see that R must satisfy the relations
I y + Sy = R(Iy + Sy)R- 1
and
(l;l:SZ + IzS;l:) = -R(lzSz + IzSz)R- 1 We automatically satisfy (8.32a) if we make R be a rotation about the satisfy (8.32b), the rotation must be 1r/2:
(8.32a) (8.32b) y~axis.
To
(8.33)
(8.23a) a result which follows from the relationships (Table 7.1) such as
which, utilizing Table 7.1, gives (8.23b) Now, we wish to apply a pulse which produces some rotalion R which will cause an echo. The effect of R is to change e from its value at t = r- 10 a new value at t = r+ given by (8.24) We want e(r+) to be such that during the next interval r, e returns to its value at t = 0+, just after the first pulse. Thus, the condition for an echo is e(2r) = g(O+) but e(2r) = e- ial• S ... e(r+)e+io1• S , ..
(8.25) (8.26)
If the effect of R on e(r-) were equivalent to changing the sign of a during the interval between the X pulse and R, so that Rg(r-)R-1 = e+ io1,S, .. e(0+)e- ia1 • S... (8.27)
then we would satisfy (8.25) and produce an echo, since putting (8.27) in (8.26) we would have e(2r) = e- ioJ• S e(r+)e+ ia1 ,S,..
Y(1r!2)I y y- I (7r!2) = I y Y(1r/2)I;l:y- 1(71/2) = +Iz Y(1r/2)lzY(7rI2) "" -Iy
and
(8.34) (8.35)
Before leaving the solid echo, we should remark on one more matter. In Fig. 8.3c we show the spins 1, 2, 3, and 4 at a time ar!2 = 1r/2. We want to refocus the spins to their condition in Fig. 8.3b where 1 and 3 lie along the +y-direction. Looking at Fig. 8.3c we note that if we simply reverse vectors 2 and 4, vectors 1 and 3 will reverse their precession directions, and thus refocus as desired a rime r later along the +y-direction. All that is needed is a 7r pulse about the x-axis. So why do we use a Y(7r/2) pulse instead of an X(7r) pulse? Clearly the idea will work if Gr!2 = 7r/2. But to give a true echo, the scheme must work for other values of a, hence of aT. The easiest test is to look at (8.32). An X(1r) pulse will leave I z and S;l: alone, but will reverse I y • S", I l , Sz. It will therefore satisfy (8.32b), but it willllot satisfy (8.32a). Note, however, in (8.29) that when (ar/2) = 7r/2, the coefficient of the lenn (Iy + Sy) vanishes, so we will refocus this particular product ar. We will not, however, refocus for either a general value of a or of r. Thus an X(7r) pulse will not produce a true echo.
= e- iol• S
Re(O+)R-1e+io1,S,T = e-iol.S, .. e+ial. 5... e(O+) e -iol. 5... e+ io1• S ... = ,,(0+)
378
(8.28) 379
The resulting magnetiUltion is shown in Fig. 8.5c. So far, everything we have done is similar to our first steps in discussing the echo using Fig. 8.3. Examination of the four magnetization vectors in Fig. 8.5c shows that their resultant is zero. We appear to have lost the magnetic order. However, that cannot be true since we know that a Y(71/2) pulse will enable us to recover the magnetizalion. The secret of the Jeeller-Broekaen pulse sequence is the recognition that the spins in Fig. 8.5c have spin-spin order. This can be understood by analyzing what happens if one applies a Y(1I"/4) pulse, producing the result of Fig. 8.5d. If now 380
Fig. 8.Sa-e. The Jeener-Broekaert method of creating dipolar order from pairs of sl)ins (I and 2, 3 and 1). (a) The thermal equilibrium magneti~ation at ! = 0- showing the vector sum of the four magneti~ation vectors. Spins I and 3 are the blue spins, spins 2 and 4 the red Sl)ins. (b) The magnetization vectors at t 0+ immediately following the X(1r/2) pulse. (c) The magnetization vectors at time t T- ,just befoN! the Y(1r /4) pulse, for the case (JT/2 = 1r/2(aT = 1r). (d) The magnetizatiOll vedors at t = T+, immediately after the Y(1r/4) pulse. (e) The magnetization vc<:tors a long time after! = T, The transverse components of mllglletization have decayed to zero from spin-spin couplings to Qlher pairs, leavmg only the resultant componenls along the ~·direction which are 1/V2 shorter than the vectors of (d)
= =
381
we have in our ensemble of spins a weak coupling of one pair with another, there will be a slight spread to the precession frequencies ± aI2. Thus, at a much later time. the transverse components of the magnetization will decay, leading to the situation of Fig. 8.Se. The remarkable poim about this state is thai every spin has its neighoor pointing in the opposite direction. This is to be contrasted 10 the condition al t = 0-. before the first pulse, when half the spins had a neighbor pointing parallel, half had their neighbor pointing anti parallel. Thus, the dipolar
engergy
haI~S;,
Indeed, because I~S~ is. apart from a conSlanl, the dipolar energy, we can see that the part of /? proportional to I~S~ is a measure of the dipolar order. If the dipolar system now comains some order, one should be able to "read" the order by convening it from dipolar order to ''Zeeman'' order, i.e. to net magnetization. Juner and Broekaert do this by applying yet one more Y(.,../4) pulse. which produces an echo a time tt later given by il ;; T. It is not surprising that an echo results since the Jeener-BroekaeIt sequence then of
which was initially zero, now takes on a nonzero value as in
an adiabalic demagnetization experiment. We employ the operator approach using a derivation due to Wang et aI. [7.73]. The density matrix at time T after the X(1f!l) pulse but before any other pulse is given by (8.23b)
(;'(,) ;; A[(Iy + Sy)cos(a,fl) -
(1zS~
+ I~Sz)2 sin(aTfl)]
(8.37)
The pulse Y(1I"/4) then rOtates the spin veClOrs 10 give
e{r+) = AY(>/4)(ly + 5 y)Y-' (>/4)
can be seen, in the limit T _ 0, to be X(rm ... r
J2
y
J2
J2
2(1~S~
- IzS z ) sin(aTfl)]
We now invoke the argument that couplings to more distant spins. not included in the Hamiltonian (8.11), cause the ofT-diagonal components of ~, which arise from the lenns (Iy + Sy) and IzS z in (8.38), 10 decay. Thus at a later time T we can drop the off-diagonal elements of ~, so that f!{T + T) is then
(J{T + T) ;;
-2AI~S~ sin(a;f2)
.
(8.39)
This Hamiltonian has zero net magnetization, as can be seen easily by explicit evaluation of Ot;;z,y.~ ;;Tr{Iau(,+T)}
(8.40)
However, the average dipolar energy
;; -Tr{haI~S~/?(r+T)} ;; -2AhaTr{I; sin(ar/2)} ;; -Alia 2 sin(aT/2) .
<7-l(,+T»
(8.41)
This is maximized in magnitude for (a,/2) ;; 11"/2 at an energy of (-Aha/2). This result is to be contrasted to its value at t;; 0- before the X(1I"/2) pulse:
<1{(0-»
= Tr{rlaI~S~e{O-)}
=
0
.
.
(8.45)
u(r + T+) ;; l"llr/4)( -2AI~S~)y-l11"/4) sin(aT/2)
;; -
•
;;
2A (I,
-
I.) (5, - 5.) . ( (2)
j2
-A[I~S~
-
.j2
IzS~
-
Sill
I~Sz
ar
+ IzSz ] sin(a,/2)
(8.46)
If this is acted on by 1{ for a time t I. then we get e{T
+ T + tt>
;; e-iol.S.tl e{, + r)eial.S.1 1 ;;
-A[I~S~ + IzSz - (IzS~ + SzI~)cos(atl!2) - (IyS; + Syi;)2 sin(ailfl)] sin(ar/2)
Utilizing the fact that
(8.47)
fi and S; are !' we get
e{, + T + tl) ;; -A [l~S~ + lzSz +2 Sy) Sin(at l
(IzS~ + Szl~)cos(att!2)
/2)]
sin(a,/2)
.
(8.48)
= i: sin(ar/2) sin(at l/2)Tr{(1y + Sy)2}
(8.49)
Of these terms, only the last one will contribute a signal:
;;Tr{(I++S+)~(r+T+tl)}
This signal is clearly maximized if we choose r and tl equal to each other, and both of such duration thai
;; Tr{haI~S~A(I~ + S~)} 382
-2AI~S~ sin(ar/2)
_(Iv
S;
(8.44)
it
So
J2
(8.38)
',
which is just the Powles-Mansfield dipolar echo. To show that this happens in the general case, we take /? just before the second pulse to be given by (8.39)
X sin(aT!2)
;; A[(Iy + Sy) -
T ••• Y(1I"/2)
e{r + T-);;
5 )_ 2A ((I. + 1.)(5, - 5.) + (I, - 1.)(5. + 5,»)
y
y(r/4)Y(>/4)
;; X(1I"fl)
- AY(7r/4)2(IzS~ + I~Sz)sin(ar/2)y-t(7r/4) = A(I +
(8.43)
x(>m ... r ... Y(r/4) ... T ... Y(r/4) ... ',
(8.42) 383
arfl ::: attfl = 7C(2
•
(8.50)
Then
'H.::: --yhhoI~ - -yhH) I~ + L Bij(3I~iI~j -1; .lj)
Comparing this signal with that of the free induction decay following an X(7C(2) pulse,
= T,{(r+ + S+)AUv + Sv» = iA n{(Iv + Sv)2}
(8.54b)
i>j
We now define an effective field in the rotating frame
(8.51)
Herr:::iH, +kh o
we see the maximum signal is one half that of a pure X(1rfl) pulse. If there is a distribution of values of a, there is still a signal at tl ::: r since (8.52) sin 2(ar!2) > 0
(8.55)
We shall concern ourselves with cases in which Herr is much larger than the local fields. In Ihal c~se, it is appropriate to quantize the spins along the effective field. Defining this direction to be the Z-direction, we make a coordinate transfonnation (Fig. 8.6) for all spins, getting
Thus, we can create dipolar order by applying X(7C!2) ... T ... Y('ll14) with T chosen such that a T ~ 7Cfl. and wecan inspect the dipolar order by later applying a Y(7C/4) pulse to produce an echo a time T later. The inspection pulse can be used to study the decay of the dipolar order resulting from either conventional spinlauice relaxation mechanisms (e.g. coupli.ng to conduction electrons in a metal), or from motional effects which modulale the strength of the dipolar coupling. We have treated the problem for coupled pairs. JCClicr and Brockacrt treat the general dipolar case in their classic paper [8.12].
Iz ::: I",cos8 + I",sin8 Ix :::I~cos8-Iysin8
Iy = Iv
•
(8.56.)
giving the inverse transfonnations for an individual spin j
Izj::: I Zj cos 8 - I Xj sin 9 IZj::: Ixjcos8+Izjsin8 Iyj::: IYj
(8.56b)
Subsriruting into 'H., we gel
8.7 The Magic Angle in the Rotating FrameThe Lee-Goldburg Experiment
+2
'H.::: -"fhHcrrIZ +
L
A/l1(8)'H/l1
where
(8.57)
/1,1:::-2
Another imponant concept concerning dipolar order is connected with the famous Lee·Goldburg experiment (8.13] 10 which we now turn. Suppose one has a number of spins all with the same precession frequency, coupled together by dipolar coupling. We keep just the secular tenos of the dipolar coupling, giving a Hamiltonian in the rotating frame of
'H.::: -
- "fllH II~ 2 "f2h + :L -,-(1 - 3 cos28ij)(3I~i l~j - Ii' Ij)
'H.o:::
,
"fhhoI~
i>j
(8.53.)
where as usual
r
h,
etc.
ho::: Ho -wl-y
(8.53b)
It is convenient to collect all the radial and angular lenns of (8.53a) in a simple symbol by defining I "f2h2 B;j=zT(I-3cos2 8ij)
Ii' Ij)
(8.58)
z
----
T ij
I~::: r:,I~j.
L Bij(3I~iIzj i>j
,
I I I I HrH ! I I H,
, x
Fig. 8.6. .The erre<:t;,·e Held in the rotAtin« rl1lme, showing the AXes X,Z WIth respect to ~,z
(8.54a)
OJ
384
38S
1-£+1 =
L
3BijUt IZj + IZirt}
i>j
1-£+2 =
L
3BijUt It}
,
i>j
1LI
= (1-£+1)", 1-£-2 = (1-£+2r
and
(8.59)
-'0(9) = !(3cos 28-1) (8.60)
-' ± 1(8} = -! sin 9cos 8 -'±2(8) = sin 28
-i
The 1-£/I.1's satisfy a commutation relation
lIz, 1iM]- M1iM
(8.61)
as is easy to show by explicit calculations of matrix elements. From (8.61), or by examination of the explicit form of the 'HM's, we see that only 1-£0 commutes with the Zeeman interactions of the effective field. In the limit of large effective field, we can then as a first approximation drop all terms but the term involving 1-£0. We get then a truncated Hamiltonian 1-£ = -yhHefflz + !(3 cos 28 - I) Bij(3IziIZj - Ii' Ij}. (8.62)
L
i>j We can compare this with the secular part of the Hamiltonian of the spins in the lab frame,1tIl\b' in the absence of an alternating field, H,: 'H'l"b = -"'('iHolz +
L
Bij(3IziI:j - Ii' Ij) (8.63)
Equations (8.62) and (8.63) are identical in form, except in the rotating frame the dipolar term has been multiplied by the factor 3,os 29 _ I 2
(8.64)
['This is juSt the tenn -'0(8)]. 8 is delennined by the relative size of ho, the amount one is off resonance, with HI, the strength of the rotating field. Lee and Goldburg noted that if they chose HI and ho properly, they could make cos 28 (8.65)
=!
in which case the dipolar term in (8.62) vanished. In this manner they could effectively eliminate the dipolar broadening. An effective field for which (8.65) is salisfied is said to be at the magic allgle. We can rephrase their result by saying that if the effective field is at the magic angle, a spin will precess in the rotating frame solely under the influence of Neff, without suffering a dephasing and consequent decay of the components 386
(8.66) The failure of the dashed theoretical curve to intercept the origin arises because of slight inhomogeneity of H1. BarntUJl and Low [8.14] made an extensive study of the systems in which HI was exactly at resonance. In this case 3cos29 - I (8.67) 2 They did experiments in which they turned on HI for a time T, then observed the free induction decay a time t after H 1 is turned off. They solved the problem of an interacting pair exactly. For an HI along the x-axis they found My(t) = M(O)
i> j
:= 1-£Zeemlln + 'H'dipolllr
of magnetization perpendicular to Herr as usually occurs when there is dipolar coupling. To make a precise test of this concept, they did pulse experiments in which they observed the decay of the magnetization with time following a sudden tumon of H t at a frequency w somewhat off resonance. They varied the angle 8 and the strength of the effective field. To measure the effective dipolar strength, they determined the second moment from the transfonn of the decay curves. They studied the F ID resonance of CaF2. Figure 8.7a shows their measurements of the square root of the second moment versus (3cos 28- l)n.. AI the magic angle, the second moment should vanish. It does not quite do so owing to the nonsecular terms involving 'H± I and H±2. The effect of these tenns should vanish in the limit of infinite Herr. In Fig.8.7b we show their measurements of the faJl-off of second moment, nonnalized to its value in the lab frame, versus llw; where
~ sin(flT)cos [ (3:,: 2 ) (t + T!2)]
(8.68)
where
n' -_ (3BI2)' - - +w,,
"
4~
(8.69)
In the limit of large HI, this expression agrees with the result of the truncated Hamiltonian: My(t) = M(O) sin(wl T) cos [ (3:,:2 ) (t + T!2)]
(8.70)
A striking feature of (8.68) is Ihal the oscillation after the tum-off of HI is identical to what it would be following an X(7r/2) pulse except for (1) a slighl amplitude correction and (2) a change of T in the apparent zero of time t. In a beautiful set of experiments, Barnaaf and Lowe showed that for CaS04 ·2H20, CaF2, and ice the first zero crossing of the free induction decay moves 10 later times as T is increased. The delay in zero crossing is T/2. for values of T which are up to about half of the nonnal free induction decay zero crossing.
387
I
(oF,IU doptlll H.1I111lJ
.. ~ 0.6
>J A
."
0.'
v
0.'
~
"-
s
A
."
., ~
pulse a time T later as suddenly lransfonning the Hamiltonian to be the negative of the actual Hamiltonian. Then, over the next time interval T, the time development of the magnetization unwinds, returning the magnetization to its value just after the initial pulse. In discussing spin temperature (Chap. 6), we talked about irreversible processes connected with the complexity of a system of many coupled dipoles. We now tum 10 a remarkable discovery which shows how one can run backwards the dephasing produced by dipolar coupling, thus showing that it is possible, even after the free induction decay is over, to recover the initial magnetization. This experiment shows that the spin temperature approach does not in all cases accurately describe the evolution of a spin system. In the process an echo is fanned. This sort of echo has become known as a magic echo. The first intimation that one could refocus dephasing arising from dipolar coupling was noticed by Rhim and Kessemeier [8.15,16], who discovered the effect experimentally, and showed theoretically by an approximate method Ihat an echo was fanned in which the loss of signal owing to dipolar dephasing was recovered. Following this work, done at the University of North Carolina, Rhim went to MlT where, in collaboration wilh Pines and Waugh, he extended and perfected the experimental techniques and the theoretical analysis [8.17]. In order to refocus spins which are defocused by dipolar coupling, one needs 10 be able [0 reverse the sign of the dipolar Hamiltonian. Then, as we have seen, we can effectively run the system backwards in time to undo Ihe dephasing. The trick is closely related to Ihe Lee-Goldburg experiment. The critical equation is (8.62), the truncated Hamiltonian in the rotating frame:
H, : 5.1!O.2 Gauss
/"
0.'
/:
'"
0.2 0.1
/'
~OO~5;:;-0::;.';:;-0:::J~-0:::.2~-0:;.1---'~0'--;0:;.1--:0::.2:-:0::J:-:0~.'---'0:':5,---J0.' ]cos'e~l
lal
2
• ';
-v., , -.,
.10. 1
//
,
/
•
/ /'//
A
~
/
/
2
A
/
•/0
//
~
v
/
/
/
(oF1lU doped I
Ho lll110l 2 /3(05 2
&.,/
~
0.01
11. = -",lill wIz + ,ell
0:'0--;:-;-----;;7'-----:':c--.J 0.1
Ibl
0.2
0.3
0.4
Il&.ll'lwli....:
F1g.8.7a,b. The ~ Ilnd ?oldburs experimental results for F" relIOnance in CaF, sins1e ct)'st.I~. (a~ The normllh~ second moment. of the Fit resonance for flo par.. llel to the ~lll J dued,.on, as a function of (3 COlI" - 1)/2, where B specifiCil the orientation of H
c:r
III the rotatmg rr.~ (see Fig. 8.6). (b) The normalized second moment iLS II function the .second moment III the lab frame < (4....') > LAn divided by",,' where .... = 11 e' e " elf110 1$ parallel to the crystal (1101 direetion
8.8 Magic Echoes In Secrs. 2.10.and 5.~ we saw thai one way of understanding how a spin echo comes ~bo~t IS to view me first pulse as initialing a time development of the
(3cos 26-1)
,,,
'"' B·1)·(3[z1z· I J) -1·I - I·) )
2~.
(8.62)
In this equation, () is the angle between the static field Ho and the effective field, Heff, and the Z-direction is the direction of the effeclive field. If Heff is nearly parallel to Ho,cos8 = 1 and the angular factor, (3cos 2() - l)Il, in front of the dipolar tenn is + I. However, if () = as when H t is exactly at resonance, the angular factor becomes 3cos 2(} _ 1 I (8.71) 2 =-1:
trn,
Thus, it has a negative sign. The magic echo makes use of Ihis negative sign to unwind the dipolar dephasing. Let us then rewrite (8.62), the truncated Hamiltonian, exactly at resonance, using the coordinates x, y, z in which HI lies along the x-axis: 1 1i = -"'thHII~ - 2" BjJ.:(3J~jI~k - I j .IJ,;) (8.72) ;>k
L
In addition to Ihese tenns, there is the nonsecular tenn
magneUZ3tJon under the influence of the Hamiltonian, and to view the second 388
389
1{nonsec =
~
L
3D jk
(lt rt + I j- Ii:)
,
(8.73)
j> k
which we are for the moment neglecting. How can we utilize the negative sign to unwind dipolar dephasing? For the moment, consider an experiment in which we produce some transverse magnetization [for example, by a Y(nl2) pUlse]. Let it dephase under the action of the dipolar system, for a time T, then turn on H I to produce a negative dipolar coupling. Will this refocus the dipolar dephasing? If the density matrix (in the rotating frame for our whole discussion) is initially (t = 0-) given by (8.74) where A is some constant which we set equal to I for convenience, the pulse sequence would produce at a time t a density matrix e(t)=ex p [
L
~1{zzT) Y(-Ir/2)Iz {inv}
(8.76) Bjk(3IzjIzk - Ij' h)
j>k
e{r+2T
)
=ex p [
x exp
-* (r!lHII~-~1{XZ)TI] [-
X exp ( =exp
X exp ( 390
*(-~!lHI[~ *
-
~1{~z)
*
2T
T ]
(8.77)
(8.80)
where tl is the observation period. Picking T ' = r will cause the signal at tl = 0 to correspond to the full magnetization. It is easy to extend the discussion to show Ihat if one were to hold T I fixed and reduce T, the dipolar refocusing would occur at a time tl given by
tt +r=T'
(8.81)
so that if r goes to zero, there would be a dipolar echo at t I = r ' . Note that if T goes to zero, the Y(n!2) and y-I(lT!2) pulses just undo one another, so both could be omitted. This is exactly what Rhim and Kessemeier did in their first experiments. From (8.78), it is clear that
= ex p ( I
')]
1iZZ T) Y(-Ir!2)Iz {inv}
(8.79)
Clearly when T I = T, the dipolar dephasing has vanished. Our pulse sequence is thus (reading left to right)
y- 1 (n/2) ... r,/!]
1{zzr) Y(-Ir/2)[z {inv}
[-1ii( - 2'it..
1{;z2T')] y- I (1T/2)
Y( n /2) ... T... Y -I( n /2) ... TIIl I ... TI-III ... Y(lTl2) ... tl
This expression has two dipolar temlS with opposite signs in front, but one is 1{~~, the other 1{zz, so they are not the negatives of each other. Moreover, there is also a term involving Ih which induces precession around the x-axis, whereas the initial dipolar term involves only 1{zz, the dipolar dephasing. It is easy to get rid of the precession, as was shown by Solomoll [8.18] in his paper describing the rotary echo. He showed that he could get rid of dephasing from precessing in an inhomogenous HI by suddenly reversing the phase of HI so that the spins precess in the opposite sense. We employ the same idea here. We keep HI on for a time TI then reverse it for an equal time r'. Then we get I
*(_
(8.75)
Bjk(3I~jI~k - Ij' h)
L
I e{T+2T ) = Y(1T!2)exp [_
=exp(*1i zz T')ex p ( -*1izzT)Y(lT/2)Iz{inV}
j>k
1{zz:=
Therefore, we add two pulses, a Y(1T/2) and a y-I (1T!2), where the latter is simply a Y( -1T/2) before and after the time interval 2r' giving
xex p ( -*1i zz r)Y(1T!2)Iz {inV}
where {inv} is the inverse of the operators to the left of I z and where we define 1{~~ ==
(8.78)
-i( -'''HII<-~'it..)(t-T)]
exp ( -
X
\
This expression is still not quite what we want since it still has two different dipolar Hamiltonians. However, examination of (8.76) shows that 1{u; and 1i zz differ simply by a cOQrdinate rotation of z into x or x into z. Such a transfoooation is produced by a rotation about the y-axis. In fact
... T I -
-*(-1t
I ZZ T
»)
Ifl
...
Y(1T!2) (8.82)
Rhim, Pines, and Waugh label such a sequence a "burst". In fact, they argue that effects of the nonsecular teoo (8.73) can be reduced if one makes a single burst by putting together many pairs of (HI,T') and (-HI,T I ) using very short TiS instead of using a single HI, - HI pair of longer duration. Their argument, the details of which they omit, appears 10 be related to their thinking about so-called average Hamiltonian theory. Of course, the most obviolls way of reducing errors from neglect of the nonsecular teoo is 10 make HI very large. Rhim, Pines and Waugh used an HI of 100 Gauss for CaF 2 with T'S of the order of 1.25 p,s. 391
In their work, they also implied that one should pick ,HIT = mr where n is an integer, although in fact they demonstrate data which violates this condition. Takegoshi and McDowell [8.19] have shown experimentally that this is evidently not necessary. It is worth noting that there is a strong experimental reason for reversing HI many times at short intervals rather than once at a long interval. The requirement of applying a large HI for a long time often places a great strain on the power supply of the rf power amplifier, causing the amplilUde of HI to droop with time. For a "burst" involving only one phase reversal of the rf, mere may be significandy lower HI during the second half (the phase reversed period) of the burst Ihan was present in the first half. If, however, there are many short cycles (H .. -HI), the fractional difference between HI and -Ht will be greatly reduced. It is interesting to note that if one has a group of nuclei with several chemical shifts the shifts can be revealed by a magic echo. During the time the HI is on, the chemical shift fields, which lie along the z-axis, can be neglected since they are perpendicular to the much larger HI. During the time T when HI is off, the full chemical shift acts. Since that is only one-third. of the time, the precession frequency is displaced from w by one-third of the chemical shift frequency. Rhim, Pines, and Waugh demonstrated that one can apply a train of magic echoes which, similar to a Carr-Purcell sequence, will refocus the magnetization again and again.
In Sect. 3.4 we described experiments by Andrew and Eades on the use of line width studies to reveal the presence of rapid molecular motions. This effect had actually been discovered by GUlowsky and Poke [8.20). For the case of rapid motion, Gutowsky and Pake had shown tht the angular factor (1 - 3COS 20jk) should be replaced by its time average as discussed in Chap. 3. That average is given by (3.61)
, , (3COS 2"Yjk - 1) 2
(1 - 3cos 8jd&Y. = (I - 3cos 8)
(3.61)
where (J' is the angle the molecular rotation axis makes with the static magnetic field, and "fjk is the angle made by the internuclear vector rij with the rotation axis. This expression suggested independently to Lowe [8.21] and to Andrew et al. [8.22] that one could produce the rotation artificially by turning the entire sample. In that case, the angle fJ' would be the same for all pairs of nuclei throughout the sample. Then if one chose fI to satisfy the condition
I - 3 co"9' : 0 392
z=rcos8
,
x=rsin8cos9
,
y=rsin8sin9
.
(8.84)
We recall that me Yim's are related 10 solutions of Laplace's equation
v'(h,m)= 0
(8.85)
[The fact that the unnormalized column of r1Yim's satisfies (8.85) is readily verified by expressing '\12 in rectangular coordinates). This is the reason that, for example, 1=2 functions are made up of linear combinations of x 2 , y2, %2, xy, %%, yz, and do not include terms such as x or xy2. Table 8.1. Listing of the normali:ted spheriul hllrmonks (or f = 0, I, 2 and the uoorlllalized forms
8.9 Magic Angle Spinning
,
the time averaged dipolar coupling would vanish [8.21, 22b]. This value of 8' soon became known as the magic angle, and such a method of line narrowing is called magic angle spinning (abbreviated as MAS) or magic angle sample spinning (MASS). It is historically the first of the methods for narrowing dipolar broadened lines. The effect of spinning on dipolar coupling is only one of several important uses of spinning to eliminate unwanted couplings. It can also be used to eliminate chemical shift anisotropies and first order quadrupole splillings. These are all interactions which involve angular functions made up of the 1 = 2 spherical harmonics, Y'm' To analyze these various methods, it is convenient to begin with a millOr digression about spherical harmonics. Table 8.1 lists the normalized Yim's for I = 0, I, and 2 as well as unnormalized forms involving the coordinates x, y, % where
Normalized spherical harmonies Yo
-
0.0 -
, "7f;
l"l,dB"P) =
constan~
-fl; sin Bel.
# cos (B, 4» = #
Y1,0(B,4» =
Yl,-l
Unnonnatizcd r'Y'm in rectangular coordinates
VIm
Y2,2(B,4» = Y2,1 (B, 4» =
Y2,o(8, 4» =
z+iy
,
B
z- iy
sin Be-i.
a --/* a(3
sin' Be,l. sin 8
COlI
Bel.
+ iy)
=z.: + iy.:
cos 2 8 - 1)
Y2._1(B,4» = V*sin B cos Be-I. Y2._ 2(8,q,) =
.:(ot
j"ji; sin 2 Be- 21 •
.:(z - iy)
=:u -
(ot _ iy? =
:>:2 _
iy.: y2 - 2izy
(8.83) 393
'.
Suppose now we consider two coordinate systems x, y, z and x', y, Zl, with corresponding angles 0, tP and 8', ¢I defined by (8.84). The Yim(O, tP)'s form a complete set for a given I, which means that we can always express }'i fi (8', tP') in tenns of the Yim(O, ¢)'s: }'jfi(e', tP') =
Lafim l'im(O, tP)
and
(8.86a)
m
Y,m(B,~) = L:em"Y,,(B',?,)
,
lal
(8.86b)
It is then slTaightforward, utilizing the onhogonality propenies of the l'im's when integrated over 4lf solid angle, to show that
x'
z'
".
I
Ibl f f / l / /
//
f y/
'. 1(1 f
I
I
Idl f
'.
L: Y,;.(e;., ~o)Y,m(B', ~')
Yo /'
/'\
I
(8.89)
01
m
(This theorem can be derived by making an expansion of a 5-function lying along the z-axis. One utilizes (8.86), the fact that the 5-function is axially symmetric about the z-axis, and equates the expansion in the lI",(D.?)'s to that in the
Yim(8', tP')'s] . We now wish to consider a spinning sample. To do this. we wish to define some axes (Fig. 8.8). We define first the laboratory z-axis. ZL, along the static field, and an axis zit about which we will eventually rolate the sample. The angle between ZL and Zit is 00 (Fig.8.8a). We can then, without loss in generality, define the laboratory axis XL to be perpendicular to both zL and Zit (Fig. 8.8b). We define XIt to be coincident with x I, (Fig.8.8c). The axes xn.. !JR, and Zit are fixed in the laboratory frame (Fig. 8.8d). We then define axes xs, YS, and zs which are fixed in the sample with zs coincident with Zit at all times, and Xs and YS coincidem with Xlt and Yit respectively at time t = 0, but making an angle nt at" later times (Fig. 8.8e). n is the angular velocity of sample spinning. Utilizing these definitions, we can specify the orientation of Ho in the coordinate system Zit' YIt, Zit (at angle 00R, 4toRl and in the system zs' !IS' zs (at angle Dos, ¢os)·
/
"
(8.88)
They can also be found using the formalism of the Wigner rotation matrices (see for example the text [8.23]). Let us now specify the orientation of the z-axis as being at an angle (9'0' ¢o) in the primed system. Then there is a famous theorem, the addition theorem for spherical harmonics, which tells us that
/
/
/
"
'. "
= b31 x + b32Y + b33Z
Y,o(O,O)Y,o(B,~) =
'.
/' /
)//
=b II X+bI2!J+b 13 Z
y' = b:ll x + b22!J + b23Z
'.
"
'.
'. "
(8.87)
Noting the forms of the Yi", 's in rectangular coordinates, we realize that the coefficients all'" or Cnll' could be found by explicit substitution of the relationship between the primed and the unprimed coordinates:
"
I
'. --
I
J../
/
Yo y,
'. '."
Flg.8.8a-e. Axes imporh.nt ror sample spinning. (a) The laboratory z-axis, ZL, is chO$en along the static field /lo. The axis ZR makes an angle 80. Together. ZL and ZR define a. plane. (b) The lab axis %L is defined as lying perpendICular to the (ZL, ZR) plane. (e) The axis %R ill coincident with %L. (d) The axes lIL And YR lie in the (:R,zd plane. Af{ the axes %L.1IL,rt. and %R,YR,zR are fixed in t~e laboratory rrame. (e) The axes %s,VS. and Zs fixed In the sampl~ and rotate with il, Zs coincidins chosen to be coincident with ZL With ZR. At t = 0,Z5 and ZR, so that it makes an angle Ot with them at later times
" u:e " "
I'
From the figure we see °OR=OO,
Oos = nOR = 00 '
(8.90,) ¢os =
tPOIt -
nt = 7rn - nt
(8.9Ob)
We now consider the interaction between a pair of spins j, k which we have previou.sl~ w~tten as proportional to 3cOS 2 0jk - I (3.7). We now, however. must dlstlOgUish between coordinate systems. hence we replace (Jjk by OLjk> writing the coupling 2 3COS 0Ljk - 1 = JI67r/5Y20«(JLjk, tPLjk) (8.91) Then, we can utilize (8.89) to get Y20(O, O)Y20(OLjk, ¢I,jk) = LY2"m(00s, ¢OS)Y2m(OSjk> tPsjk)
(8.92)
m
395
where 0Sjk, ¢Sjk give Ihe orienlation of the internuclear vector Tk in Ihe coordinate system xS' Ys. zs which is fixed in the sample and where Oos and ¢os are given by (8.90). Substituling the explicit expressions for the Ylm'S from Table 8.1, we get
1
3 COS 20Ljk _
= (3COS
2~OS -
1)
We turn now to a discussion of the effect of spinning on a pulse experiment in which we observe the free induction decay of a spinning sample following a -;r/2 pulse. Then we might express the dipolllr Hamiltonian, 11.d as
(3cOS20Sjk _ I)
j
(8.93)
We see that the second and third tenns on the right vary as f}t and 2f}t respectively, whereas the first tenn on the right is independent of time. Thus if one can average the time-dependent tenns, we recover the result of (3.61) with the notation
Cs = 3 CO!;200S - I
(8.95)
2 the/orm of the absorption line should be identical in the two cases except for a scaling of the frequency. Thus, if i(w - wo) is the nonnalized intensity function without spinning, Ihe normalized function with spinning is(w - wo) becomes fs(w - wo) ,
1
C/(Cs (w - wo))
FOjk
=
Fljdt)
=
for large 'Y nuclei, since spinning speeds are typically a few kilohertz, comparable 10 typical dipolar line widths.
(8.98b)
(3COS200S - 1)(3
2
'" COSVSik- l )
6 sin Oos cos 80S sin 0Sjk cos 0Sik COS(¢Sjk - ¢os)
(8.99)
F 2jk (t) = ~ sin200s sin 0Sjk coS 2(¢Sjk - ¢OS) Since ¢os = -;r/2 - f}t from (8.90b), we see that FOjk is independent of time, but F1jk and F2j~' oscillate in lime at f} and 2f} respectively. Explicil examination of the spin functions G jk shows Ihal two different funClions which have one spin in common, for example the jth, such as Gjk and Gjl (f of- k), do not commute. This makes it difficult in general to determine the time development. Such mailers are discussed in Appendix J dealing with time-dependent Hamiltonians. Here, we are dealing with Ihe case of a Hamilitonian composed of a sum of noncom muting parts each of which has a lime-dependent coefficient. This is Case II of Appendix 1. However, the lime dependence is periodic, so we can consider what happens over a single period. Suppose that the period T(::= 21f/f}) is sufficently shan that not much happens during one cycle. Then, as we show in Appendix I, we can relate the wave function at the start of a period to its value at the end of Ihat period !>y the unitary operator U(T) given by
,p(T)' U(T),p(O), exp ( - i7tT) ,p(O)
(8.l00)
where
l'
-11. T'fo =
(8.97)
3Izj l zk )
2
(8.96)
At the magic angle, the line theoretically becomes a O-function. In practice, it is difficult to achieve the condition
-
is a function only of Ihe XL, YL, and ZL components of the spin operators Ij and Tk' and
~~
For steady-state experiments (i.e. if one talks about the "frequency" domain, perhaps with data which are the Fourier transform of data from a pulse experiment), the tenns involving f}t and 2f}t give rise to frequency modulation at frequencies f} and 2f}. Such a modulation gives rise to multiple sidebands spaced in frequency by f} aboul the ncnnal fequency wo. For frequencies f} much bigger than the dipolar line width ~Wd' these sidebands will be displaced by an amount large compared to the width associated with the time-independem tenn. If we neglect the sidebands to focus our attention solely on the main transition, it will appear narrower. Indeed, since the time-independent tenn of the dipolar coupling for the spinning case is simply reduced from its value for the nonspinning case by the faclOr Cs where
396
where 2 ·2 o" Gjk(Tj , h)::= -,-(Ii' h
r jk
(8.90b)
~k=~k
(8.98a)
j>k
Since ¢os obeys (8.90b), 10s ' ,/2 - fit
1 jk
, I: [FOj' + F1j'(') + F'j,(.)jGj,(Ij,I,)
+ 6 sin OOS cos Oos sin 0Sjk cos 0Sjk cos (¢Sjk - ¢os) + ~ sin200s sin 20Sjk cos 2(¢Sjk - ¢os)
...,2h2 3 COS 20Sjk)-.-3-(3Izjlzk - I j ·Id
E (l -
'H<J =
But
(8.lOl)
'H(r)dr
T
T
jF1jk(r)dr=0
jF2ik (r)dr=O
o
0
so that it is only the term FOjk which brings about a change in period.
(8.102)
t/J over an
integral
397
This is tantamount to saying that to the extent F1jk(t) and F 2jk(t) produce changes during the time interval T, they must also undo these changes over a complete period T. The resuh is that the free induction decay has periodic maxima which Lowe called rotational echoes. Figure 8.9 shows the data from Lowe's paper [8.21] for FI9 in Teflon. The rotational echoes are clearly apparent, in Fig. 8.9a. as is the fact that at the magic angle (54.7°), the free induction decay lasts much longer. The Fourier transfonn of these data (Fig. 8.9b) shows the sideband structure. It illustrates the fact that even when spinning at over 6 kHz, the sidebands are still not separated from the central line for the high FI9 nucleus.
,,---,----.----,---,-----,
•
F (I)
(aJ
,_
• , •
'H,
Is flle/s) 0.3.1,6..2 _ _
".1
"
".1
"
Andrew et al. [8.22] and Kesscmeier and Norberg [8.24] demonstrated clear separations of the spinning sidebands for lower 'Y nuclei Na 23 and p31 respectively. hKieed, removal of the dipolar coupling between like spins is done more readily by applying strong pulses to flip the spins, so-called spin-flip narrowing, which we discuss in the next section. We can understand physically the significance of the requirement (8.97) that n>6.wd and its relation to the fact that G ij and Gjl do not commute as follows. For spinning to average out an interaction, the interaction must be constant in time over at least one cycle of rotation. To average the coupling GjJ:' the local field that spin j produces at spin k must not change during one cycle. But the coupling Gjl pennils spins j and I to undergo mutual flips, thereby interrupting the coherent averaging of G jk' 6.wd is a measure of Tjl' how long spin i can maintain its orientation without such a mutual spin flip (Tjl ~ 1/6.wd ). For narrowing of the jk coupling to occur, the period of rotarion (1/0) must be less than Tjl, Tjl>
02
"
1/n where
Tjt"" I/.6.wd
(8.103.) (8.103b)
These can be rewriuen to give (8.97)
3 2
{»
'" (b)
..
I (I's/
""
f!
,I
3
P
"
"" ii'G~
2
f"'
__'_ r'W t F(11
G(WI
201684048121620
W( rod/s) x 10'4
1<'1g.8,9, Free induction decays (8) for spinning and nonspinning samples of Teflon, and their Fourier lransforma (b). The curves are corrected for instrument.l nonlinearities. The data are due lo t. L/)wt, 1/. Ktsst~itT. W. Ytn. G. Thriss, and R.E. Norbug 18.201
398
.6.wd
(8.97)
TIle case we have just discussed, which we call Case II in Appendix J, has been called an example of a "homogeneous" broadening mechanism by Maricq and \Vaugh (8.25]. (This tenn is closely related to an idea introduced by Portis [8.26] in electron spin resonance). Its essential feature is that Ihe time-dependent parts of the Hamillonian do not commute with one another. To produce line narrowing in this case, one needs n ;» .6.w where 6.w is the line width from this mechanism. In the absence of spinning 6.w = 6.wd for dipolar broadening. We now tum to a case which has a much less stringent condition, and thus is much easier to achieve experimentally, narrowing lines when the time-dependent parts of the Hamiltonian commute, Case I of Appendix J. We illustrate this case by line broadening from the anisotropy of chemical or Knight shifts. This case is called "inhomogeneous" since it is analogolls to line broadening by inhomogeneous magnetic fields. Andrew recognized that magic angle spinning would remove line broadening from sources, other than dipolar coupling, which depend on the orientation of Ho with respect to the crystal axes. The key point was that the broadening be described by appropriate angular factors arising from the orientation of the static magnetic field, Bo, with respect to the crystal axes. When the angular factors are proportional 10 Yim's with 1 = 2, there is a magic angle. Other cases are line broadening from chemical and Knight shift anisotropies, and from electric quadrupole splitlings. We treat the case of shift anisotropy.
399
Recalling (4.198), we get that, in lhe presence of anisotropic shifts,
Y2,0(9 0p, ,pop)
=
L Coms(P, S)Y2ms (90S• tPos)
'H. = - 'Y;,Ho[1 + (/( _ U)+(/(LO _ 0LO) (3COS:O- I) +
(I(TR;
UTn) sin 2ocos2 tP] l z
.
(8.104)
We consider lz to be the sum of the z components l z k of N noninleracting spins (k = I toN). We shall analyze the effect of rotation on this Hamiltonian. The angles 0 and tP of (8.104) specify the orientation of Ho in the principal axis system of the shift tensors. We shall need to keep track of various reference frames. Therefore, we shall define !he spherical angles Oop and tPop to designale the orientation of fIo in the principal axis system xp, yp, zp. We keep the same definitions for the coordinate systems XL, YL, zL (Ho lies along ZL), %R, YR, Zn, and %s, YS, Zs (Fig. 8.8). We have then that at t = 0, ZL is at Oop, tPop or expressed in R or S at 00R, 4>oR or OOS. 4>oS. which in tum are described by (8.9Oa). Therefore. we can wrile at t = 0
Y2 _2(9 0p , tPoP)
,
At later limes. owing to the rotation, fos will change with time. bUI since the rotation axis, 90S does not change, so 80 5(')
=80
(8.105)
As time goes on, Oop and t/Jop change, owing to the sample rotation. hence should be viewed as time dependent. Recognizing that the two angular functions are related to Y2rn'S apart from nonnalization conslanlS. we define lhe nonnalizations N 2,o, N2,2' N2,_2 as Y2,0 Y2,2 Y2,-2
== N2,0(3cos 29 - I) == N2,2 sin gei .p == N 2,-2 sin ge- i .p
,
=.f2 -
ill
zs is
(8.109)
.
Therefore the mS = ±2 lenns of (8.108) will oscillate at ±2n, the ms = ± 1 terms will oscillate al ± and the ms :: 0 tenns will be time independent. Collecting lerms we gel
n.
'H.= -'YhHo1z[I+(J(-U>
+ .
(8.108)
+ ( 1(1.0 - 0'1.0) "L...COms(P,S)Y2rns(8o.'ll"f2 - fI) t 2N2,0 ms
'H.= -'YhHO[1 +(l( -U)+ (J(LO ;ULO) (3 cos 290p -I) + (KTn;UTR) sin20oPcoS24>OP]lz
no, = L C-2Jl1s(P, S)Y2ms(Oos. tPos) no,
(8.106)
and express
"L...[C2ms(P,S)+C-2,ms(P,S)I ( [{TR4N-2,2OTR) ms X
[Y,no,(8 0,.f2-flT)I]
(8.110)
This Hamiltonian consists of several time-dependent parts all muhiplied by the same spin funclion, l z • Thus the Hamiltonian alone time, t, commutes with the Hamiltonian at another time, t':
['11('), '11(,')) = 0
(8.111)
.
This HamillOnian therefore belongs to Class I of Appendix 1. Indeed, it is even described by (l.28) (1.28)
'II(t) = a(t)'II.
for which the formal solution is given in (1.30) as ,p(') = exp ( - T.'II.
We can now express the angular functions in tenns of the S-system, using (8.86),
400
f
a(T)dT) ,p(0)
.
(1.30)
Let us first consider just the time-independent term (mS = 0). It will give a contribution to the Hamiltonian of
401
1tlimeindcp= --rhHQTZ{ 1 +(J( -01+ [ +
(/(Tn -
(J(L~N~,:I.O) CO,O(P,S)
a TO ) [C20(P,S) + C_, O(P,S)I]
4N2,2
x N2,0(3cos 280
'
,
1£ = - ,hHol, {I + (I( - 'i7) + AI(P.S)cos
-I)}
--yfiHolz[I + (I( - a)J
.
(8.113)
£
7{(T)dT = -'YhHo1,
C',o(P,S)
= a;,,(S,P)
Y2~O(O.O)
Y;,(a,fIj • aO,2 = yi,o(O,O)
'!>(') = exp ( -
Y2.,O(O,O)
(8.115)
where (a, fJ) gives the orientation in spherical coordinates of the spinning axis
[(I(LO-O'LO) (3Ca"a-l) 2 {1+(I(-'i7)+ P)] (3CO"Bo-I)} ' _O'TR)(,inaCOS2 + (} 2 2 .
1£timeindep= ~'YfIHo[,
(8.116)
(This expression can be checked in the limiting case that 80 = O. hence the spin axis lies along Ho. Then the spinning has no effect and the time-independent tenn is the entire Hamiltonian. Then 0' = 0 or 0' = rr(2, {J = 0 or 1r(2 correspond to.flo lying along the three principal directions. The result is that for the three cases we get shifts respectively 1(3 - 0'3. 1<1 - 0' 1, or J( 2 - 0'2, where 3, I, and 2 stand for the zp. x p , and Up directions.) The other very interesting point to consider is the time dependence when one is spinning at the magic angle. Examination of (8.110) shows that the time dependence of 1£ comes from the factors Y2ms (80, 1r(2 - nt). Ulilizing the prop-
402
(8.118)
*£
7{(T)dT) ,!>(O)
(8.119)
and recalling that T = 27f/n. we see that when
Y;,_,(a,p) °0,-2::
%s in the Xp. yp. %p (principal axis) coordinate system. The result is
\TR
"1JI
Utilizing the fact that
and the spherical hannonic addition theorem (8.89)
Y;o(a,p)
+(I( -
. [ + A,(P,S) 2il ( sm 2ilt - r2(p,S)] + sin r 2(p,S)} )
(8.114)
C_',o(P,S) = a;,_,(S,P)
ao,o =
([1
. [ + AI(P,S)( n smm-r 1 (P,S)]+sinrl (p,S)}
(8.87) = a;,o(S,P)
(8.117)
where AI(P.S) and A2(P.S) are coefficients, and rl(p,S) and r2(P.S) are phase angles, all of which in general will depend on the orientation of Zs in the xp, yp, zp system. Utilizing this fonn, we get
Thus, it is identical to the fonn taken in a liquid in which the anisotropic shift contributions average to zero. If 80 is nO[ at the magic angle we might wish to know the coefficients CO,o(P, S), C2,O(P, S), and C-2,O(P, S). These we get from the inverse relalions
CO,o(P,S)
[nt - rl(P.S)]
+ A,(P,S)co, [W. - r,(p,S)))
(8.112)
The coefficients C 2,0 (P, S) and C- 2 ,0 (P, S) can be obtained (if desired) as we explain below. But even before that, we nOie thai when 80. the angle between the rotation axis and Ho. is 3t the magic angle. we gel 'Htime indep =
erties of the Y2ms 's (Table 8.1) we see that 1£ is a sum of tenns which are either independent of time. or vary as exp(int), exp(-inl). exp(2int). exp(-2int). Keeping in mind that the Hamiltonian must be real. and that we are at the magic angle,.we can thus write
t=nT
•
n=O,I,2 ...
'!>(I) = exp ( - i'YHo1,(1 + [( -
(8.120.)
a)') ,!>(O)
(8.12Ob)
just the result we would have if the time-dependent terms were missing. Note that this result says that at the magic angle the time development over integral multiples of the rotation period is independent of the orientation of the crystal axes relative to Ho. Thus, if we applied a 7f{l. pulse to a spinning sample. the free induction decay at times given by (8.120a) would act as though there were no anisotropy to the shift tensor. If one had a powder sample which was not spinning, the spread in precession frequency arising from shift anisotropy. 6.wshif~' would cause the free induction to decay in a time 1/6.wshif~' If, however, the sample is spinning we can see from (8.120b) that the signals are rephased at times given by (8.120a). thereby producing a string of echoes. Notice that we have not placed any requirement as yet on how fast the spinning must be, in contrast to the previous siwalion concerning narrowing of dipolar line broadening. For the dipolar case we compared the rigid lauice line width ~d with n. What frequency should we compare with w in the present case? Dimensionally, there is only one relevant frequency. the total excursion
403
of the precession frequency over the spinning cycle. Note that all our equations so far imply there is a single crystal being rotated since there is a well-defined orientation of zs in the xp, yp, zp coordinate system. However, for a powder, all possible orientations occur. For the rest of our discussion we will focus on a powder. Then, the maximum frequency excursion one can have in a rotation cycle is less than 1'Ho[(J(zI'Zp - qzpzl') - [(I{:cI':C1' - q:cp:cp)]
== .6wshifL
n '"
404
•
(8.121)
where we are defining the zp and xp axes as having the maximum and minimum shifts respectively. So we ask, what happens when we vary the relative sizes of fl and .6wshift.? Suppose, first, that the powder sample is not spinning. Then, as we have noted, following a rr/2 pulse, the transverse magnetization will decay to zero in a time :::::: l16.wshift. The Fourier transform of this free induction decay would give us the line shape. If we then applied a 11" pulse at time tll" we could refocus the magnetization into an echo. If t1l':» 1/.0.wshifL' the echo and the free induction decay would be well separated in time. Indeed, we could apply a string of such 11" pulses (a Carr-Purcell sequence) pnxlucing a string of echoes. The Fourier transform of anyone echo will give us the powder line shape in frequency space. If, instead, we took a Fourier transform of the sIring of echoes, we would now have introduced a periodicity. The resultant J1ansform differs from that of a single echo in the same way that a Fourier series differs from a Fourier integral. The transform of the string of echoes would consist of 5-functions spaced apart in frequency by the angular frequency l/t rep where t rep is the time between successive rr pulses. The Fourier transform of a single echo would be the envelope function of the spikes. If now one shortened t rep , one would get to a point where t rep :::::: lI.6wshifL' Then the echo would not have decayed completely to zero when the 11" pulse is applied, so that over a single period one no longer would have the complete shape of a single free induction decay. The Fourier transform of the sequence of echoes would still have a center line plus sidebands spaced in angular frequency by 1/t rep , but the envelope of these sidebands would not be the powdeI line shape (the rransform of a single free induction decay). Stejskal et al. [8.27] realized that the same situation would arise from magic angle spinning. When n« .6wshifL, the NMR signal following an initial 11"/2 pulse decays rapidly compared to the period of rotation, hence is essentially identical to what it would be with O. However, after one full rotation there is an echo, as described by (8.120). Indeed, a train of "spinning echoes" will be formed, analogous to a Carr~Purcell train. The Fourier transform of any echo will give the 0 powder line shape. The Fourier transform of the train of echoes will give o-function spikes spaced apart in angular frequency by with an intensity envelope versus frequency identical to the nonspinning powder patlern. This situation is illustrated beautifully by Fig. 8.10 showing the data of Herzfeld et al. for a p13 compound [8.28,29].
n '"
Proton decoupled p3l spectra at 119.05 MHz of barium diethyJ phosphate spinning aL the magic angle. The figure illustrates how, in the limit of slow spinning, the spinning sidebands reproduce the shape or the Ilonspinning spectrum [8.28] .'ig.8.IO.
I'AQT- OkHz
n,
"OOT - 0.94 kHz b
I'OOT"
2.92 kHz
d
16
,
a
0
-
a
, -16
Freq (kHz)
n
Once becomes comparable to or larger than .6wshifL> the satellites occur outside the envelope of the nonspinning line. As can be seen fIOm (8.117), the coefficients A\(P,S) and A2(P,S) ale multiplied by lin, hence become progressively smaller than the lalger n. This is part of a general theorem discussed by Andrew et al. [8.22] that the second moment is invariant under rotation. Thus, the intensity of lines spaced by multiples of must fall off as l/n 2 in the limit of large n. When there are many spinning sidebands, how can one tell the central line? The easiest technique is to change n. One line will be undisplaced - the cenrral line.
n
405
The beauty of the slow spinning regime of Stejskal, Schaefer. and McKay is that it gives one simultaneously a high precision determination of the isotropic average shift, as well as a picture of the associated powder paHem. It also removes the difficulty of spinning at a rate faster than ~Wlihirt., a genuine difficulty in the strong fields of superconducting magnets when one has large shift anisotTOpies. A common use of the slow spinning regime is to slUdy low abundance nuclei (e.g. C 13 ). so that there is no like-nuclei dipolar coupling with which to contend, combined with standard decoupling to any high abundance nuclei like HI whose mutual spin-flips would otherwise broaden the spinning side bands. In some instances it is desirable to remove sidebands in order to simplify the spectra. This can be done by applying 1( pulses synchronously with the rotation as shown by Dixon [8.30) and by Raleigh et al. [8.31].
H( 1 r[5 ) =
406
S
M[ - -3-
r lS
r lS
(8.122)
We consider M[ to lie along the k-direction, and consider three cases in which S lies at a distance a respectively along the X-, y-, and z·axes (see Fig.8.ll). Application of the formula shows that when S is on the z·axis it experiences a field 2M, HI(O, 0, a) • - 3 (8.123) a whereas when S is on the x· or y-axes HI(a,O,O)' HI(O,a,O)' - M.,I (8.124) a If, for some reason, S were to jump rapidly among the three positions, spending equaJ limes in each on the average, it would experience a time-averaged field < H/ > given by
8.10 The Relation of Spin.F1ip Narrowing to Motional Narrowing The strong spin-spin coupling characteristic of solids was essential for achieving the high sensitivity of double resonance. Its disadvantage is that it produces line broadening, and thus may obscure details of the resonance line such as the existence of anisotropic couplings (chemical shift, Knight shift, or quadrupole interactions). Within the past few years, a varielY of clever techniques have been introduced which utilize strong rf pulses to eliminate much of the dipolar broadening. We now lum to the principles on which the ideas are based. The highly ingenious concepts owe their development to several groups. The pioneering experimenlS were performed by two groups, one headed by John Waugh [8.4], Ihe other headed by Peter MallSfitld [8.S]. Subsequently Robert Vaughan [8.6] and his colleagues have contributed imponantly, as have many of the scientists [8.32,33] who worked with Waugh, Mansfield or Vaughan. The essence of these schemes is to apply a repetitive sel of rf pulses [8.3439) which produce large spin rotations and which, by a process akin to motional narrowing, cause the dipolar coupling to average to zero. Each cycle consists of a small number of pulses (8 pulses per cycle is common). To achieve the effect, Ihere need to be many cycles within the normal dephasing time of Ihe rigid lattice line width. Since the rf pulses must produce large spin rotations (90" pulses are typically used), and since the rotations should occur within a small fraction of a cycle of pulses, one needs H ,'s which are large compared to the rigid lauice Ijne breadth. We have remarked on the similarity between the multiple pulse schemes and motional narrowing. We stan by making the analogy explicit. For convenience we shall name the multiple pulse schemes "spin-flip narrowing". Consider two spins, I and S. Let 1" [5 be the vector from spin I to S. The magnetic field H[ at spin S due to I is then
3(M/-rIS)r/5
(8.125)
, ,..---... , ,
,,
,,
,
/
, s (,j
,
, I
x
s
"F----
(b)
s
\, ,I
, ,,
IF----
(c)
Ag. 8.lla-e. The magnetic field or [ III S for lhree localions of S all the nme distance, a, from ,pin I. The dashed line indicates II. magnetic line of force
407
This result is the essence of motional narrowing. Though for our case we picked only discrete locations for spin S, the same result is found for continuously variable sites for which the strength of the interaction goes with angular position 81 s of rlS with respect to the z-axis as 3cos 2 81S - I. This averages to zero over a sphere. Suppose now we consider a variation in the above picture. Suppose we position S at (0,0, a), Le. on the z-axis, and consider the orientation of both M I and Ms (Fig. 8.12). In tenns of the orientation of the vector r I S with respect to MI' it is useful to introduce some names. Referring to Fig. 8.11, we denote the (c) configuration as the on-axis position of S, we denote the position of S in (a) and (b) as the side position. Note that for the on-axis position HIS is parallel to MI, whereas for the side position HIS is anti-parallel to MI and half of its on-axis magnitude. Referring now to Fig. 8.12, we see that in (a) and (b) S occupies a side position, whereas in (c) it occupies an on-axis position. Moreover, since we have taken M , and Ms parallel in all parts, for configurations (a) and (b) the magnetic energy E llIag = -Ms' H,s is:
a quick rotation of both spins to the side position of Fig. 8.12b. Wait another time 'T. What is the average magnetic energy < Em > over the interval 3'T? Using (8.126) and (8.127) we get
- .!...(-2MI MS'T MIMs'T ll'hMST)_o + + a3 a3 a3
< E m> - 3T
.
(8.128)
That is, over such a cycle in which, at stated times, we apply selected 7fn. rotations to both spins, we can make the magnetic energy average to zero. Thus by spin-flips we can cause the dipolar coupling to vanish just as we could for position jumps in Fig. 8.11. This is the principle of spin-flip line narrowing. We make the dipolar energy vanish by flipping the spins among selected on-axis and side positions, spending twice as long in the side positions as in the on-axis ones. All of the complicated pulse cycles are based on exactly this principle.
8.11 The Formal Description of Spin-Flip Narrowing
(8.126) whereas for (c) Emag)e =
-Ms' HIS =
-2NhMs
(8.127)
a3
Suppose, then, we began with the configuration of Fig. 8.12c, the on-axis arrangement with MI and Ms parallel. After time 'T let us quickly rotate M I and Ms by 7fn. about the y-axis to the side position of Fig. 8.12a. After a 'T give
,
(,)
S
~x
, (
(b)
S
(c)
S
"
" (
FIg.8.I2. The two spill!l Ai, and M. are oriented parallel to one another. In (a), (b), lind (c) their magnetic moments are parallelrespcctive!.y to the Z-, Ih and z-directiolls
408
From the previous section we see that it is possible to cause the dipolar interaction 10 average 10 zero if we cause the spins to flip between the "on-axis" and "side" arrangements. For the conventional motional narrowing, narrowing occurs when the correlation time Tc and the rigid lauice line breadth (in frequency) DwRL satisfy TcDwRL $ I
(8.129)
Withoul jumping, a set of spins precessing in phase initially get out of step in a lime ..... (l/OwRL)' The condition expresses the faci that for motional narrowing to occur, the jumping must occur before the dephasing can take place. In our example of motional narrowing, the correlation time would be on the order of the mean time spent in one of the three configurations of Fig.8.11. If T is the time in anyone orientation, roughly Tc
:::::: 3T
.
(8.130)
In a similar way we expect that spin-flip narrowing will worle only if the spins are flipped among the needed configurations before dephasing has occurred. Thus we get a condition on T:
(8.131) The better the inequality is satisfied, the longer the spins will precess in phase. We expect, therefore, that we shall wish to flip the spins again and again through lhe configuration of Fig. 8.12. This we can do by applying a given cycle of spin-flips repetitively. The basic cycle will bring Ihe spins back to their starting point at the end of each cycle. 409
A fonnal description of what takes place begins with a specification of the Hamiltonian. We shall use it to compute the development of the wave function in time. We write lhe Hamiltonian as 1i(t) = 1irr{O + 1iint
where
(8.133) (8.134)
1t3 = 2: Bij(rjj)(Ii .Ij -
(8.135)
'Ho includes the chemical and Knight shifts qUi."H~ is the secular part of the dipolar coupling, the coefficients Bij(f"ij) including the distance and angular factors [see (8.54a»). The objecl ofspin-jlip narrowing is to cause ~ to vanish while maintaining 1to nonzero. To illustrate the principles of spin-flip narrowing, it is useful to idealize the situation. Correclions to the idealization are important as a practical matter. We return to them in Sec!. 8.13. The idealization is to consider that 11rf is zero except for very short times during which it is so large that 'Hint can be neglected in comparison. This approximation enables us to say that 'Hrr produces spin rotations at the time a pulse is applied, but that between pulses the wave function ¢(t) develops under the action of the time independent Hamiltonian 'Hint. Thus between pulses the wave function at lime t can be related to its value at an earlier lime t I by (2.49) ¢(t) = exp{ - (ilh)'Hint(t - tt»)¢(t t)
(8.136)
which defines Uillt. The effect of pulse of amplitude HI and duration t w along the a-axis (a = x, y, z) at time tt is to produce a transfonnation by the unitary operator Pj for the ith pulse, 410
Pi = e i(1rj2)lo
(8.138)
if the pulse produces a 1r/2. rotation about the a(= x, y,or z) axis. Let us consider a three-pulse cycle for simplicity and concreteness. We always begin an experiment by having a sample which has reached thermal equilibrium in the magnet. We then tilt the magnetization into the :t - Y plane at t = O. We call that pulse the preparation pulse. After a time TO we begin applying the repetitive pulse cycles. Choice of TO and the phase of the preparation pulse is made in tcnns of producing a signal at some point in the pulse cycle which is convenient for observation (at a time which is called "the window" in the spin-flip literature). Let .p(tn)be the wave function after n cycles [Le., just before pulse PI of the (n + I)th cycle]. Then we can follow the wave function in time with Table 8.2. Table 8.2. Varialion of the wave function with time for. three-pulse sequence Time
3I::iI::j)
ij
= Uint(t - tlhb(t)
(8.131)
where, for example, "'{Htt w is chosen as 1ffl giving
(8.132)
We shall work in the rotating reference frame with the z-axis defined as the direction of the stalic field. 1irr{t) is the coupling to the applied rf pulses used to apply the spin rotations. It is time dependent because the pulses are switched on for very short intervals, two In principle the corresponding H) can be oriented along the x-, y-, or z-axes. Selection of the x- versus the y-axis is a question of the phase of the rf pulse. Though pulses along the z-axis can be applied in principle, in practice they are not used since they would require addilion of ,mother coil 10 the rig. However, a z-axis rotation can be achieved by two successive rotations about the x- and y-axes. (Show how this is done!) The tenn 1tin~ consists of two tenns, 1tint = "Ho+~
¢(tt) = ei..,lI,tw l o .p(ti)
I; (just before PI) (just after ~)
It t.
+ T,-
(just before 1'\)
+ Tt (just .ner P,) I. + Tl + Ti" (just before 1\)
,,= "('.)
.p=P,¢(I,,)
'" =exp( -i1ti•• 1"1 j")~ ¢(t.. ) =UI.,(TdPt.p(I.)
t.
¢ =
I.. +TI +T1+T3-
tP = Ui•• (T1)1'\ Ul •• (Tl) PI ,,(t.) ¢(I,,+d = UihtCr~)/"JUi ... (T1)P:!Ulnl(Tl)Pl\/.,(t,,) == UT ¢(! .. ) defining UT
Dust before (II
+ l)th cycle]
~Ui .. (TdP,
tP(t.)
We can thus write (8.139) where UT is independent of n. Since UT is a product of unitary operators, it is itself unitary. After N pulses, we have
,p(tN) = u!f,p(to)
(8.140)
so that the problem can be considered solved if the effect of UT can be deduced. Let us examine one cycle, and introduce the unitaty operators p j - I which are the inverses of the Pi'S. For a unitary operator Pi-I = P/
(8.141)
where the" stands for complex conjugate. 411
Then we write UT as UT :::: Uill~(T3)P.JUin~(T2)P2Uin~(Tt )Pt :::: P3 P2 PdP\-t p 2- t P3- 1Uinl(T3)P3P2 PIHP.-l p 2- t UinL(T2)P2Ptl
X [P1-tUinl(T)Pd
(8.142)
Now we showed in our example of Sect. 8.10 that a complele cycle of spinflips should get us back to the starting point so that we can repetitively flip the spins among "on-axis" and "side" configurations. Therefore the cycle PI, P2, PJ should get us back where we started. Hence (8.143) giving us
(8.144a)
:: Uinl(TJ)[PI- 1P2- 1Uinl(T2)P2PI][PI-IUinL(TI)Pl]
(8.144b)
The meaning of the individual tenus can be made evident by several transformations. Erst, consider a unitary operator P and a Hamiltonian 1t, and the corre· sponding U: P-IU(t - to)P:: p-l e -(i/l)1{(t- 1o) P
.
(8.146)
so that the effeci of p- 1 UP is 10 cause U to develop in time under a transfonned Hamiltonian. If we were to evaluate the expression in the exponent of Ihe transfonned Uin~(T2)'
Pit P2-I1tinL(T2)P2Pj
(8.147)
we could first do the trunsfonnation p2-t'HinlP2, then sandwich the result between p t- 1 and PI and evaluate that, using the appropriate expressions for exponential operatOrs. 111is order of application of the opemtors is the rever.fe of the order in time of application of Ihe spin rQlation operators. Why is that? Consider a SchrOdinger equation
_':.iN=1i'" (8.148) i 'f' We can transfonn this equation with a unitary operator R, which is independent of time, to the problem
at
412
(8.150)
as the 7t corresponding to the rQlation P- I • We can consider the following two descriptions of the effect of 1t~ following 3. 1fn. rotation: i) Use ~ untransfonned acting on a
t/J which
is rotated
+wn.
ii) Leave the spin function alone bUI rolate the spin coordinates in ~ by
-wn., the inverse of lhe spin rotation in (i). Equation (8.144) corresponds to (ii).
(8.145)
By expanding the exponential, insening p-l PC:::: 1) between factors, and regrouping, we find that p-I U(t - to)P = e.p[ - (Uh)(P-'1iPXt - to»)
which leaves the problem lhe .fame. Thus, if R were a spin rotation operator giving a +1r{2 rotarion of the spins about some axis, R1tR- 1 musr transfonn the coordinates of Ihe Hamiltonian corresponding to the same rotation. If 1t:: -..,hHoIt and the spin is in a Slate corresponding to the spin-up state, a transformation, R, which rotates Ihe spin function into the up spin state along the +y-direclion requires rotating 1t Ihe same way, which is done by replacing I: by I" in 1t. We can therefore interpret p- l 1tP:: (P- I )1t(p- I )-1
UT :::: (P1- 1P2- 1P;'Uinl(T3)P:! P2P j][P1- 1P2-IUinL(T2)P2Pd X (PI-IUinl(TI)Pd
(8.149)
If a rolalion is made up of several rolations in succession, the inverse consists of lhe inverse rotations perfonned in the opposite order. Thus, if the spin is flipped by P2P\, the inverse transfonnmion Q is (P2PI)-I, so the 1t transformed by the inverse is
(8.151) The prescription for finding the transfonned 1t conesponding to the ith interval Ti of an n pulse sequence is to take the spin rotations PI, P2 Pi, which preceded the interval and transform the coordinates in the Hamiltonian by applying rhe inverse rotations in the reverse sequence (e.g., first p;-t, then Pj-=.ll , ••••... , lastly PI-I). We now define the three rransfonned Hamiltonians 11.", 'HB, and 'He as 1t" :::: PI-I1tintPI 1tB :: P l- l P2-I1tinLP2PI
(8.152)
11.e :::: 1tint On expanding the exponentials we get
413
('H~)B : 'L Bij(Ii ·Ij - 3lzi l zj )
UT =ex p ( -*1iCT3)eXp( -*1iBT2)eXp(-*1i"'11) = 1-
i<j
i
(~>c
1i (1iCTJ+1iB1'2+1-£",Tt)
• L: B(I; . I j -
3I,;I,j)
(8.158)
i<j
and
- (*Y(1iCT31iB1'2 + 1tC7'J1-£", 11 +1iB1'21iA11)
T) :
7"2 : TJ.
(8.159)
2 2 +'H. 2 7"22 + 'H.",7"I) 22 + ... +-l(i)2 - (1-£C7"3 B 2 I,
(8.153)
If the 7"'S are short, (8.154) where by 1(I'HA,B,clhl)1 we mean a quantity of the magnitude of typical matrix elements of the transformed Hamiltonians. Under these circumstances, UT is well approximated by keeping only the leading two teons on the right of (8.153). The condition on 7"i is similar to the requirement on the correlation time 7" for there to be motional narrowing. Introducing the period of a cycle. t c = Tl + 7"2 +"f'J, we get
(T'J
7"2 7") ) UT = I - -i 1ic- +'H.B-+'H.Atc II tc te te
i __ = I -1i'Hinttc ';:!
ex p ( -*'H.inttc)
(8.160)
Setting w : wo for simplicity. we would then get
Ho :
liwo"" -3 ~ azzi(Izi + I y ; + lzi)
,
(8.155b) (8.161) (8.155c)
'H.in~ : -h(wo - w)Iz - liwQ 'Laui1zi + 'H~ (8.156)
The trick then is to choose the pulse cycle (PI> P2, P3, etc.) so that we eliminate the dipolar coupling but maintain Ihe chemical shift and Knight shift information: (8.157.)
-h(wQ - w)lz - hWQ 'Lazzi1zii-Q
(1io )", : -h(wo -w)III-llwo 'Lazzi1yi
(8.155.)
where (b) defines 'H.in~' the average 1iinh and (c) serves to remind us that over a cycle the system develops to a good approximation as though 1iint were replaced by its average 1iint' defined above. Now we expand. 1iin~ into its elements
" 110 + 11~
These results could be achieved if p,-I were equivalem to a 1r/2 rotation about the x-axis which would transform Izi imo [yi and if P 1- 1p-I were a 1r12 rotation about the y-axis which would transform Izi and Izi. 2 These pulses would also transform 'Ho
(8.157b)
(8.162) Equation (8.162) shows that the chemical shift and Knight shift are reduced by v'3 (i.e., multiplied by 1/Jj) from their value without averaging. We have succeeded in making the dipolar coupling vanish without eliminating the chemical shift. To the eXlent that the term in (w - wo) is included (it represems frequency offset or field inhomogeneity). it is also reduced in the same proportion. According to (8.139) and (8.155c), ¢(to) and f/J(t) are given by
,,(t) . ul( ,,(to) : exp( - (i/li)1tinl(t - to»)¢(to)
(8.163)
provided
t = to + Nt c
(8.164)
How abour ¢(t) at other times? Consider a time t in the Nih interval such rhar
Thus if
(~)A : 'L Bij(li .Ij - 3111iIlIj)
t:tO+Ntc+tl
(8.165)
i<j 414
415
The time t) could fall into anyone of the time intervals T), T2 • ••• Tj which make up the basic cycle t c . Over anyone complete cycle there is no large change in ¢ if (8.152) is true. However, each large pulse produces a sudden big spin change _ typically a 1fo, rotation. These large pulses might, for example. progressively cycle the spin along the X-. y-. and z-axis in the rotating frame. If one always observes during the ith interval. however. the effect of the big pulses in successive cycles will always have returned the nuclear magnetization to the same direction in the rotating frame. Thus we can write ,,(t) = e
'0»)"«0)
(8.166)
when t and to are both within the same subinterval Tj of tc' If they are within different intervals Tj and Tj. with (8.167)
Ntc
we should properly go back to expressions like (8.136) and (8.137). 1f we consider thm the individual T'S are so short that (8.154) is true, then 'Hint does not produce much of a change during t e , but the pulses i + 1, i +2, i + 3, ...• j - I, j do. Thus, defining P
== PjPj-I ...• Pj+2Pj+l
(8.168)
there are two expressions which are nearly equal:
,
pexp( - (i/h)1{intNte)tP(to) ';!
add either a steady-state or a pulsed spectrometer to observe the character of the resulting resonance. In practice. the amplifiers of most steady-state apparatus would be blocked by a large rf pulse at their input, with a recovery from blocking which would be much slower than the short time T between spin-flipping pulses. Consequently. to see spin-flip narrowing one goes to NMR apparatus designed to handle intense rf pulses-one uses a pulse rig. With coherent pulse apparatus. one picks out a particular component (Mz or My) of the magnetization in the rotating frame. If one is adjusted to detect M z • then one finds in general that it is largest in one panicular interval Ti of the cycle. perhaps zero in others. One could start with the system in thermal equilibrium. apply a 7(0, pulse about the x-axis to rotate the magnetization to the % - Y plane. and then start the spin-flip cycles. For simplicity of the discussion we assume we can apply 7ro, rOlations about the %-. y-. and z-axes in the spin-flip cycle. though in practice z-axis rotations are not used. A possible set of pulses to initiate and carry on spin-flip narrowing is shown in Fig. 8.13. In Fig.8.13a we see M along the direction of the slatic magnetic field (the +z-direction). AI t = 0, Po rolates M to the +y-direction. This is called the preparing pulse. At time TO we initiate the spin-flip pulses Pl. P2, P3 which successively rotate M by -rr/2 about the z-axis, +rr/2 about the y-axis. and lastly +2rr/3 about the -i + j + k axis. The rotation P3 is deduced as being the inverse of P2P) and can also be thought of simply as PI-I P2- 1 mathematically. or physically as the successive steps -7I:n. about the y-axis followed by +71:/2 about the z-axis. Note
P exp ( - (i/h)7-{int(N + I)tc) ,p(to)
(x)
,, ,,
.......- - - , Po
(8.169)
Since Nt c < t - to < (N + I)t c we can write
.p(') '" Pe
'0»)"('0)
(8.170)
relating ,p(to) in the ith interval to ¢(t) in the jth interval of a pulse. Equation (8.167) says in essence to compute the wave function evolution between t and to as though 1{inl acted the whole time, and was followed by the rotations Pj+ I, ....., Pj-I. Pj in quick succession.
x
,
!-i+J+.t
---
(d)
8.12 Observation of the Spin-Flip Narrowing How can one experimentally observe the effect of spin-flip narrowing? Ordinary motional narrowing can be seen either by steady-state or pulsed apparatus as very narrow lines in the former case. and as long Bloch decays or slowly decaying echoes in the laller. To produce spin-flip narrowing large rf pulses must be applied to the sample. Assuming one has the equipment to apply such pulses. one could then in principle
,
i
' -, P,
,, ,
axis
i i .......... ----...., I'J
i
\
416
(b,
,
i
,, ,,
(l')
,
Fig.8.13a·e. The erred of ~he various pulses 8S llCCll in the rota~ing frame. (a) Thc initial thermal equilibrium configuration wi~h lhe ml
417
that ahhough P J looks superficially like a rOlation of -1(/2 about the x-axis. it is not. as can be verified if one considers rotating a three-dimensional object in PI, P2. P3 instead of just a vector M pointing initially along the y-axis. This distinction is imponant because lld is a second rank tensor, thus it should be thought of as an ellipsoid rather than as an arrow. If the apparatus were adjusted to measure M z • and if it were tuned exactly to resonance. including any possible chemical and Knight shifts. there would only be a nuclear resonance signal between pulses PI and P2. If there were a resonance offset (as with a second class of nuclei chemically shifted from the first), M z would gradually change in the PI-to-P2 interval for successive pulses. undergoing an oscillation (see Fig.8.14) as we show below. It would appear during the PJ-to-Pt window as well as time went on. Let us tum to a mathematical formulation of the calculation of the observed NMR signal. For concreteness, we compute the time development of M z in the rotating frame < Mz(t) >. We will use the language of wave functions rather than density matrix. Given the wave function !/J(t), we have < M,(t) > =
< ,p(t), M,,p(t) >
= '''<,p('),1,,p(r»
(8.171) We assume that we have a pulse sequence similar to that of Figs. 8.13 and 8.14. At t = 0- we let the wave function be t/J(O-); then the preparation pulse gives us ,p(0+) = Po,p(O-) (8.172)
The wave function then develops under 'Hin~. If we ask for it in the interval PJ to PI. there will be an integral number of cycles PI followed by P2 followed by PJ, so that. using (8.166).
,p(t) = e -(1M"; .. ' Po,p(O-) If we ask for
t/J(O later in
(8.173)
the cycle. we use (8.170)
,p(t) = Pe-(I/.),,;o,'Po,p(O-)
(8.174)
where P = PI or PIP2 depending on whether t lies in the interval Pt-to-P2 or P2- to-PJ. We now need to express the fact that the initial wave function 1/J(0-) cor· responds to the initial thermal equilibrium of the sample. To do so we express y,(0-) in terms of the laboratory frame 4»(O-} instead of the rotating frame. In general the rotating frame wave function t/J and laboratory function ifJ are related
by 1/J(t) =e-ilo.ltl·ifJ(t)
(8.1750)
so that at t "" 0- we get (8.175b)
Taking I n) to be the eigenstates of the secular Hamiltonian 1(. in the laboratory frame. we then have (8.176)
"
Thus <Mz(t»
=
='Yh E cnc~(m I Po-le(i/.)'Hiall P- t IzP ",m
xe-i/.)'H;a,IPo!n)
.
(8.177)
We now make a statistical average over the coefficients Cnc:" utilizing the fact that the states are occupied according to a Boltzmann distribution of the Hamiltonian 1(.. which describes the system before application of the pulses:
e- E.. lkT (n le-'HlkT I m) cnc m = Onm Z = Tr{e 'HlkT}
_._ (,)
P,
P,
p,
(8.178)
p,
Substituting (8.178) into (8.177) we get
,10 Tr( - p-llzPe-(ilh)"H,nt - t Poe- 1ilkT ) < Mz(t) > = Z Po-te(i!h)"Hinl! (8.179)
In the high temperature approximation we can write (II)
/'0
1',
1'1
I'j
P,
1'1
1',
P,
1'1
Flj;.8.14a,b. Grllph of ~he OllCilloseope signal of the resonance venus time for ~he pulse sequence when (a) apP"ra~us is se~ exactly a.~ re&Onan<:e, or (b) II chemical shifl. causes ~he M z signal to change with time appearing in ~he PJ·to-p. inter....11llI it diminishes in ~he p'-to-f\ window
.,8
e-"/kT
Z
(I -7{JkT)
"" (21+I)NG
(8.180)
where there are No spins of spin I. The leading term vanishes, so that the nonvanishing term gives a I/kT contribution. 419
If we then wrile
::: NO'Y2h2I(l + l)Ho
1£ ::: ~'YhHoIz
ttansfer the
PO-I
3'0'
(WOU"')]
from the lefl to the right in the ttace, using
1 2 :::Mo [ 3+3cOS ~
Th{ABC}: T,{BCA}
(8.182)
and if we pick Po 10 make
PolzPO- 1 = I z
(8.183)
which rotates the spin from the z·axis to the x-axis. we get
2 H {~ } <M (t»:::'...,2h °Tr e(i/1I)'Him,tp- I I Pe-(i/Il)ni·'(I z (2I + I)No kT z z· (8.184) A useful exercise for the reader is to derive (8.184) using the density matrix language. For simplicity we take P ::: I (Le., an integral number of pulse cycles PI. P2, P3 pass between t ::: 0 and the time of the measurement). Now, 1£int = -
[.!.. + ~ 3
3kT
(8.181)
m;o 2;:Uni(lzi + Izi + IIJi)
(woont)] J3 (8.189)
Mo
where is the thermal equilibrium magnetization. We see, therefore, that the chemical shifl-Knighl shift tl zz shows up as an oscillatory contribution to the < Mz(t) > ,reduced in frequency by 1/../3 from the normal frequency shift W1)Ou one would see if one could resolve il in a steady-state resonance. Jf more Ihan one kind of chemical or Knight shift is present, the resultant < Mz(t) > 's add, Ihe respective strengths being in proportion to the number of nuclei possessing that chemical shifl or Knight shift through the Mo faclOr. The form of (8.189) is that of a conStant term plus an oscillatory term. By taking a Fourier transform one can deduce a "reduced chemical shift frequency" (WOtl u/../3). II is possible to utilize Fourier transform NMR attachments therefore to decompose a transient signal when there exists man: Ihan one Knight shiftchemical shift to unscramble the discrete frequencies.
•
(8.185) where [z'i we get
=
1£inl::: -
(1/J3)(lzi + IIJi + IziJ. Suppose all spins have the same
(Tn'
hwo~zIzl
Then
(8.186)
Now in terms of the prime coordinates we have, in an obvious notation,
I z ::: Iz'cos(x, x') + 1IJlcos(x. y') + Iz'cos(x, z') Izl ::: Izcos(x, Xl) + IIJcos(y, x') + lzcos(z, x')
Our discussion up to now has been directed 10wardS explaining the principles of spin-flip narrowing. Trnnslating these notions into prnclice. we encounter certain problems. We discuss these briefly, Ihen list a few of the pulse sequences which have been developed to cope with them. An excellent discussion of these problems has been given by Haeberlen in his NATO Summer School lectures. I Vaughan and colleagues [8.38,8.39] give a thorough theoretical discussion including explicit fonnulas of use. See also the references in the Selected Bibliography.
(8.187)
8.13.1 Avoiding a z-Axis Rotation
so that exp (
')I
-iwouutIz J3
zexp
::: cos(x, x') [IzICOS
+ cos(x, y')
[-l
+ cos(x, zl)Iz'
(iwouutIzl) J3
(WOjizt) + III,sin (wojizt) ]
z 'Sin (w0./izt)
+ Iy'COS( W0.,ij-zt) ] (8.188)
Using the fact that Tr{Iz ' III'} = 0, and Tr{I;,} ::: Tr{~I} we get, after evaluating the traces, 420
8,13 Real Pulses and Sequences
Pulse rigs apply HI'S in the x - y plane. By choosing the rf phase, a single one of these can produce rolations about any arbittary axis lying in the z - y plane, but not about the z-axis. To produce a z-axis rotation with a single pulse would require adding a coil wilh an axis parallel to Ihe z-axis plus the electronics to energize it. In order to avoid so doing, one can go to a four-pulse technique. That in a sense is equivalent to composing a z-axis rotation from successive x-axis and y-axis rolat.ions. Ordinarily, however, the two pulses are spaced apan in time so that the basic timing intervals can be maintained more simply. 1 The 1974 NATO Advanced Study Inslilule held in Leuvcn, Belgium fcalur«! a number of userul lectures on NMR in solids./fa,,/)#!rl"/1 aoo Van /f«~ in p"rticulllr lreat spin-nip nflrrowing in [8.401.
421
To write sequences, it is convenient to introduce a notation based on the fact that1r(l pulses are nonnally used. For example, consider the following expression which defines one four·pulse scheme. X(-r,X, T, Y,2T, Y,"',X,T)
(8.190)
.
The first "X" represents the preparing pulse, a 1fn. pulse about the +x-axis. The parentheses enclose the pulses of a single cycle, which in the example indicated is considered to stan with a time interval T prior to application of a 1((1. pulse about the -x·axis (or a -1fn. pulse about the +x·axis). This sequence could also be written as X,T(X,T,Y,2T,Y,T,X,2T)
,
which simply chooses to call the staning point in the cycle as the time just prior to the first -:I: pulse in the chain. This notation makes explicit the presence of two intervals of 2T and twO of T. The first notation is useful in proving certain theorems which utilize the time symmetry of a single pulse cycle. 8.13.2 Nonidealily or Pulses Pulse Duration. We have utilized instantaneous rotations in the discussion up to now, which would require an infinite HI of zero duration. Actual pulses have nonzero duration. One's first guess might be that there would be major problems as a result of the nonzero duration. Theoretical investigation of this point reveals, however, that it is not serious [8.34,35,38,39]. lbe point is that slow rotations can still maintain the property of averaging out the dipolar coupling. They are in this sense equivalent in motional narrowing to allowing (3 cos 20 - I) to average to zero over a conrjmwus variation in 8 as opposed to the discrete variation of Fig. 8.1. The important experimental consequence is that much lower If power is needed that one would at first suppose. Inhomogeneous HI. As we discussed in Sect. 8.3 it is always difficult to make HI homogeneous over a sample, panicularly when one is attempting to keep the rf coil small to cut down on the power requirements (for a given Q and Hj, the power needed is proportional to coil volume). This problem is a bad one in that it causes the rotation angles to vary throughout the sample. It can be removed to first order, however, by using the trick of Mejboom and Gill or of pairing positive and negative rotations about any given axis, thereby achieving a refocusing effect. Phase Modulation. Turning on an HI which points along a given direction in the rotating frame sounds easy, but in practice is hard to achieve. Usually the phase of the HI swings during the buildup and during the decay. Such swings can be labeled as modulations of the phase angle tP of the rf. They are equivalent to being off-resonance L\w = dtPldt during the time that dtPldt=l-O [8.41). Their
effect can be eliminated to first order by having l[1>/dt during turn-on and dtPldt during turn-off equal and opposite. Making l[tPldt zero is not practical, but the balancing during turn-on and tum·off is. Among the various pulse sequences used or referred to in the literature, we shall list three. They are known as WAHUHA, HW·8, and REV-8. These mysterious symbols are made up from initials of their inventors plus numbers which tell how many pulses there are per cycle. The WAHUHA is named after J.S. Waugh, L.M. Huber, and U. Haeberlen
[8.34J. \( is X, T(T,X, T. Y,2T, Y,T,X, T)
This sequence can be sampled at the times corresponding to just before the parenthesis by measuring M lI • The name WAHUHA refers to the contents of the parentheses. The preparation pulse can be chosen to give convenient sets of timing. The panicular sequences shown have the property that the X,T, at the start are in fact like the last pulse and interval of the cycle; hence no special timing intervals are needed at the stan. That is, the preparing pulse in this case is pan of the cycle. For the preparing pulse shown, the magnetization is initially turned to the y--direction. Thus, if one detectS My at the inverval between the X and -X pulses, one will get a signal. We can write the HW-8 [8.35] as (T,X,2T,X, T, Y,2T, Y, T, Y,2T, Y,T,X, 2T,X, T)
There are a number of equivalent variations. The MREV-8 [8.38] sequence is
X, T(T, X, T, Y,2T, Y, T,X,2T,X, T, Y,2T, Y, T,X, T) This can be viewed as pair of pulses X, Y separated by T followed at 2T later by their inverse y, X. MREV demonstrate the highest resolution obtained with their MREV-8.
8.14 Analysis of and More Uses for Pulse Sequence The importance and utility of Illultipulse techniques goes far beyond spin· flip line narrowing. We turn now to several examples, as well as to illustrations of some calculations of the effects of the pulse techniques. Before beginning our discussion, it is useful to introduce a notation and 10 demonstrate some useful relationships. We shall be following the effect of pulses on individual components of spins such as I:r:, I y • or I: and on dipolar tenns. For the dipolar tenn we introduce the notation 'Hon(a = x, y. z) defined as
423
1i&& =
L
(8.191)
Bik(/j ·lk - 3I&jI&k)
i>k Explicit calculation then shows that
1iu +1iIlIl +1iu =O
~4~J
.
(8.192)
In general. the effect of pulses Pi will be rotations. To discuss the effects of nonzero pulse lenglh:i. one must treat the Pi 's as time-dependent operators such as (8.193) where t is the time variable during the pulse. For our present purposes. however. we shall Ireat the pulses as negligible duration (so-called 6-function pulses) and moreover treat them as all being 1tll pulses. We then have to evaluate expressions such as
P1- 1I: P,
(8.194a)
P,-t HuPt
(8.194b)
But once we know what (8.194a) is. it is easy to get (8.194b). For example. let
Pt = X
~.I~
using right-handed rotations. For a pair of pulses PI followed by P2. then the lefl-handed rotations
PI-tP2-t1&P2PI
(8.I99b)
would be replaced by the right-handed rotations Pt.l'2. etc. p)
P21&P2-J p)-t
(8.199<:)
Using either approach, one can work out a table for (8.198) which is lhen useful for all applications in which we list 1a across the top and in each row list first the operator PI. then the effect of Pt on 1&. The result in the table is the same whether we treat PI as a left-handed rotation and use (8.198), or a right-handed rotation and use (8.199a). Rotation operator
Operator being rotated
Iz
(8.195.)
then
Pi] = X
PI = X then P 1- J = X) and then using the left-handed rotations or by instead pretending that one is calculating
III
I:
x
then
(8.195b)
-I, Iz I,
y
Z
X1:X = - ly X1i u X =
L
or expressed as coordinates x, y.
Bij(X Ii' [/cX - 3X I:jl:/cX)
%
i>k =
L
X
Bij(lj·Ik-3XI:jXX1:kX)
y Z
i>k =
L
Bii(lj ·lk - 3IlI jIlJk)
i>k = 1iYIJ
%
z y
Y
,
y
%
,
(8.200)
2
Y
.- ,
(8.201)
(8.196)
where the bar indicates a negative. Suppose then that in the sense of (8.198) one has the operator HYII acted on by PI = X. Then the table shows us that
(8.197)
(8.202)
Thus. if
Pi] I&PI
= kl{J
Pi 1 H&&PJ
k= ±1
then
Let us now illustrate this with a calculation of the effect of a very useful = Hpp
whether k = +1 or - 1. The effect of a (Iefl-handed) rotation PJ in the expression
Pit I&PJ
(8.198)
can be worked out once one knows PI either by first computing P 1- 1 (e.g. if 424
pulse sequence utilized by Warren et al. [8.42] to refocus dipolar dephasing. They employ it as pan of their method of generating multiple quantum coherence of a particular order, a topic we discuss in the next chapter. In discussing magic echoes, we showed that one could refocus the dephasing arising from dipolar coupling by creating a dipolar coupling of negative sign. In particular we showed how we could create a dipolar Hamiltonian 1id given by 425
1{d =
-~1{n
(8.203)
We now show how Warrell et al. created such an 1{d by pulses. The basic approach is founded on (8.192) 1{z",
+ 1i yy + 1i u = 0
Thus
-11.",,,, = 1i yy + 71.:u
(8.204)
Therefore, if we create a Hamiltonian which spends equal time as 71. yy and 1{u, its average value will be }inn. Since without pulses 1id = 1i u , we simply need pulses to switch 1i u to 71.'11'11' Clearly, all we need is X's or X's for the Pi'S. The simplest thing one can do is to switch 1i u to 71.'11'11 with PI = X, then switch it back, P2 = X, so that (8.205,)
have omitted the "T' and the "1{" since only the subscript and sign is needed to specify .the entry. On the third iine we list P2, then the product P2 PI, which we simplify to the extent possible. The last line labeled "sum" contains the sum of the time intervals, and the sums of the products of the operators with the time intervals for which each operator acts. The average values are then just the weighted sums of the operators divided by the duration. Using (8.204) we see that the pulse sequence does produce an average Hamiltonian (8.206) Of course, the pulse sequence must be repeated continuously for as long as one wants to satisfy (8.206). A pulse sequence which gives the same average dipolar Hamiltonian but eliminates the chemical shift tenn and has some other advantages is given by Warren et a!. [8.42]: X,~X,~X,T,X,T,X,T,X,~X,T,X,T
for a sequence (8.205b)
(X,T,X,T)
To follow the action, we make a table. We need to follow two tenus in the Hamiltonian: the dipolar coupling and the chemical shift. Ir we simply wish to create a negative 1id for all spins without regard to their chemical shifts, we need a pulse sequence which will make all chemical shifts zero. We will find that our simple (X, T, X, T) sequence has problems with chemical shifts. We construct a table in which time progresses downwards (Table 8.3).
We evaluate it in Table 8.4 utilizing (8.201). Table 8.4. Sequence (X, T, x, T, X, T, X, T, X, T, X, T, x, T, x, T) Duration
X
, 2
Average
X
X'
X
X
(-l y +l.) Iy)
tU. -
11..
1i
yy
1i
yy
.,
"
,
"
X
yy
X
X'
X
X
.,
" yy
y
+ H .. hHJY + H.,) H
X
yy
Sum Average
In Table 8.3, the "duration" column is measured in units of
T.
8
, 0 0
4(h u
"
+ H yy )
1(H.. +HJY)
The column
PI ... Pi lists PI, PI P2, PI P2P3, etc. as in (8.199c). Remcrl"!ber, the first operator
to apply is the right hand one, but its effect is to be computed with right-handed rotations. On line I we list Pl' Then on the second line under I z and 71.u we list the operators produced by the action of Pion I z and Hz%> as well as the duration of the time between application of PI and the application of the next pulse P2. We 426
X
I,
y
yy
Sum
X
X
I,
P; P j =X
P, ... Pi
X
Table 8.l. Sequence (X,T,X,T)
Duration
P;
The same approach can be used to analyze the spin-flip line narrowing sequences discussed in Sect. 8.13. For example, in Table 8.5 we analyze the fourpulse sequence (X, T, Y, 2T, Y, r, X, 2r). As remarked in Sect.8.l3.l, this pulse sequence is equivalent to the three-pulse average we employed (rotations about the X-, y-, and z-axes) in Sect. 8.12. Utilizing Table 8.5 we see it produces a zero average dipolar Hamiltonian. 427
Table 8.5. Sequeoce (X,,., Y ,2T,Y, ",X, 2,.) Durll~ion
p.
y
J~
... P;
I,
X
,
j7
XYY = X
X
XYYX= I
•
..'"
y
yy
T.ilble 8.6. SBml)lc
"
Line
,
2
Sum Average
M..
XY
2
2(r. + J,
6
So if we knew what QIII was. we could immediately evaluate (8.212). Since fJ will be x, y, or z (or %. y, Z). we can handle any eventuality if we know the results of Qlz• QI", and QI~. A procedure which will provide us with these three results is to replace the single column I~ with three columns Iz,!",!z. A portion of the table would then look like Table 8.6.
+ Iz )
~(l.:. + J,
+ J.)
2(1l:... +1l,,+1l.. ) f(ll zz
+ 11" + 11,.)
For multiple quantum excitation, Warren et al. [8.42] show that it is useful to generate an effective operator proportional to 'H"" -'Hu . In panicular, they
I 2 3 4 S
~Ilble
to help ill
eVlllua~jng
f.
I,
f,
I~II
I"
I,
I",,,
I.,
I.
the
effcc~
or QP•• I
PJ ... p.. _ 1 PI ... P.. (= Q)
P, "'P"P"+ J
6
find
!('HYfI -'Hzz ) = i(2'H fIfI + 1l u
(8.201.)
)
is generated by the eight-pulse sequence
(x. 21'.
X.
T,
X. 21', X.
T,
X, 2,., X.
T,
X, 2,.. X,
T)
(8.208) I~
Pn+1I~ = Ipl
(8.207b).
After a bit of experience in calculating tables, one soon discovers a problem. Suppose one has just calculated the effect of the operator
acting to the right. Suppose it produces a result I a in the have to calculate the effect of the operator
To compute line 6. we first find the effect of P n + 1 on I z • I y • and I z . There will be three results. We look for them on line 2. Thus. suppose
column. We next
Then
Pj ... P,d/p = la' So we would enter I rr on line 6 under the I z column. This approach requires evaluating three columns. but it makes the step of going from one line to the next much simpler - one has only to evaluate two operator steps per column. hence six operator sfeps per line.
(8.209)
Now it would be easy if the operator we had to calculate was Pn+l Q since then (8.210)
This situation would let us take the result of Q. which we already have. and apply Pn+l to it. However. what we must calculate is instead QP..+1
which has the PII +l operate before Q operates. Therefore, we must go through all n + I operator calculations. a tedious process for long pulse squences. Now suppose Pn+lI~ =
IfJ
(8.211)
Then (8.212)
428
429
9. Multiple Quantum Coherence
9.1 Introduction In Chapter 5 we introduced the density matrix as a powerful tool for the analysis of magnetic resonance experiments, In analyzing the case of spin we saw that the diagonal elements of e were connected with the magnetization parallel to the static field, and the off-diagonal elements were related 10 the transverse components through the equations
!'
(5.257.)
11<
(M,(O) ~ 2[,++(0 - e--(O]
(5.257b)
In a system with two spins I and S we saw that
(I+(O) ~ L::(-msle(t}1 +ms}
m,
(7.248)
so that the transverse magnetization of the I-spins arises from elements of the density matrix which are off-diagonal in the I quantum numbers, but diagonal in the S quantum numbers. There are other off-diagonal elements of (2, for example
(-+le(OI+-) (- -1,,(01 ++)
0'
(9.1.) (9.1b)
It is a fundamentaltcnct of statistical mechanics that for a system in thennal equilibrium such tenns are zero, using the hypothesis of random phases described in the text just below (5.87). This hypothesis is based on two ideas. The first is that all observables of a system in thennal equilibrium must be time independent. This requirement really expresses the meaning of "equilibrium". It is a necessary (though not sufficient) condition for thermal equilibrium that all observables be time independent. The second idea is that since the time dependence of I! comes only in the off-diagonal elements, there is some experimental means for probing every offdiagonal element of I!. If there were not, the existence of off-diagonal elements would not produce time-dependent observables, to conflict with the hypothesis of equilibrium. Since we have just seen that resonance experiments do not directly 431
detect elements such as in (9.1), we naturally wonder if there is any experimental method for probing the existence of such matrix elements. A closely related question is whether there is any way to excite such matrilt elements if they are initially zero. We are already in a position to say that in cenain cases we can eltcite these elemenls because indeed we have analyzed in detail an eltperiment which generates them. Referring to our analysis in Sec!. 7.26 of the pulse sequence (9.2) we see that in (7.338) we produced tenns of form, are proponional to
{l
which, eltpressing
{l
in operator
These gave nonvanishing elements of
{l
(7.341b)
where in each matrix element the upper signs go together, and the lower signs go together. The matrix elements (7.341 a) join states of energy difference, i1E, given by (9.4)
and the matrix elements of (7.34lb) join states differing in energy by i1E::: ± 1i.(wOf - WOS)
(9.S)
If the two spins were identical, so that WOf ~ Wos, we would refer to these energy differences as two quantum and zero quantum and we would say that the pulse sequence (9.2) has generated 0- and 2-quantum matrix elements of (l in addition to the I-quantum matrix which we observe directly. A next question is, clearly, can we detect the zero- or two-quantum matrix elements given the fact that (7.248) shows that resonance directly sees only the one-quantum matrix elements? Before proceeding with that question, we discuss a bit more the concept of 0-, 1-,2-, and in general n-quantum matrix elements. In some cases, mf or ms are not individual constants of the motion. For example, consider two identical spins, with a chemical shift difference and a spin-spin coupling involving tems such as 1· S. If the spin-spin coupling is comparable in magnitude to the chemical shift difference, ml or ms are not constants of the motion. It is still true, however, that the operator, F::, for the total z-component of angular momentum (9.6)
commutes with the Zeeman energy and the secular part of the spin-spin coupling, and thus its eigenvalue, M. is a good quantum number. 432
~-E~ k:::t
~
the eigenvalue M of ['IT is a good quantum number. Thus, if we have a set of quantum numbers M and a, where a are all the other quantum numbers needed so~ , I,TIMa) = MIMa)
(9.7b)
we define matrix elements (9.8)
as n-quantum matrix elements where
n _IM-M'I (7.3410)
i1E::: ± 1i.(WOf +wos)
N
(M'a'Ie(OIMa) (9.3)
1z 5"
In fact, for an N -spin system with total z-component of spin I::T given by
(9.9)
We follow the convention of calling n the order of the matrix element, and say that the existence of a nonvanishing matrix element (9.8) describes n-Qrdtr cohertnct. It is immediately evident that for a system of N spins of spin I. M can range from N I to - N I, so that the largest n one can have is 2 N I. Thus, for protons, a spin system involving an H2 group (as in a CH2 fragment) should pennit generation of two-quantum matrix elements, a system involving three protons (as in a CH3 group) should pennil generalion of both two- and Ihree-quantum matrix elements. The generation of a given order of coherence thus enables one to determine what spin groupings exist if one can detect the existence of orders different from one. NOle, however, that a pair of CH2 groups contains four protons, so one may ask, when should one consider a CH2 fragment to be a pair of protons, and when should it be considered to be part of a larger group? We thus see that there are exciting possibilities to use multiple quantum phenomena to classify spin groups, but that there are also imponant things we need to understand, such as, how do we create and how do we detect the various orders of coherence? Multiple quantum phenomena are encountered in all branches of spectroscopy. Bodenllausen [9.1), in his excellent review of the field, includes a discussion of the manifestations of multiple quantum phenomena in cw magnetic resonance speclroscoPY. It was the advent of pulse methods in multiple quantum speClroscopy which caused the technique 10 become of central importance as an almost routine speclroscopic tool. In two fundamental papers, Hatallaka et aJ. [9.2,3] showed how to generate and to detect multiple quantum coherence. They sludied the Al 27 resonance of A1 20 3, which can produce multiple quantum transitions since the spin of Al27 is 5/2. Aue et al. [9.4]. in their famous initial paper on two-dimensional Fourier lransform NMR, show that, using two-dimensional Fourier transform 433
spectroscopy, one can observe both zero- and double-quantum speCtra. They describe several ways of producing the desired matrix elements. Since the publication of the basic papers from the laboratories of Hashi and of Emst, the field of multiquantum coherence has exploded_ 1n addition to the review by Bodenhausen. there is an extensive review by Weitekamp [9.5]. There are useful reviews by Emid [9.6J and by Mu/Wwitz and Pines [9.7]. As in aU aspects of pulsed NMR, the book by Ernst el al. [9.8) gives a thorough and detailed account of the principles and practice. Our goal is to explain the physical principles. We turn next to showing that it is the nonlinearity in the equation of motion which makes possible excitation of multiple quantum coherence. We initially deal with cases in which we are tuned to discrete transitions, showing how we can then progressively pump higher and higher order coherences. Nexi we deal with the much more common experimental case of applying pulses of sufficiently greal amplitude to excite a number of transitions, discussing as well how to detect the fact that we have generated multiple quantum coherence. We then lum 10 melhods of singling out in the detection coherence of a desired order, and lastly to the problem of how one can selectively excite a panicular order of coherence.
9.2 The Feasibility of Generating Multiple Quantum Coherence - Frequency Selective Pumping Consider for the moment a two-spin system. We have seen (Sect. 7.5) that we can produce two-quantum elements of f! by applying a driving signal at WI = wOJ, followed by one at Ws := Wos. Among the mauix elements produced is
(- - 1,,(')1 ++)
•
We now proceed to show that this result is closely related to the idea of the beating of the frequencies of two oscillators. Beats are produced by putting input signals into a nonlinear device so that the output signal contains tenns proportional to the product of the input signals. Thus if we had a device whose electrical propenies are characterized by the property that the current i through the device is proportional to the square of the voltage V applied across the device
i.AV' application of a voltage
434
j(O
:=
A(V? cos 2 Wit +
(9.12)
W2t
(9.13)
vl cos2 w;et + 2Vt V2 cos WI t cos w;et)
(9.14)
Utilizing the fact Ihal cos WI t cos W2t
!{ cos (WI + W2)t + cos (Wt - W2)t]
:=
(9.15)
we get i =
~ { V12 [1 + cos (2w 10] + vlp + cos (2wzO] + 2VI V2[ cos
(WI + wz)t + cos (WI - wz)tl }
(9.16)
w,
Thus the current i contains the frequencies 2w), 2w 2 , + W2, WI - W2, and 0 (which is WI - WI and w;e - W2). A similar thing happens for the densily matrix equation df!
d' •
i
,,(eli -1/e)
.
(9.17)
Since this equaJion involves a product of f! with 1l, if both f! and 1i are time dependem on the right hand side, they couple to tenns on the left hand side which oscillate at the sum or difference frequencies. We want to show that Ihis is just what we need to take a {} which has an n-quamum coherence and pump it either up 10 an (n + I)-quantum coherence or down to an (n - I)-quantum coherence by means of a time-dependent 11. which oscillates at the frequency difference between Ihe nand n + I (or n - I) quantum coherence. To make those ideas more concrete, let us consider a system with a set of nondegenerate energy levels, a static Hamiltonian 1iO' and a time-dependent drive tenn V(O given by V(t)/h:= Fe
(9.11)
+ V2 cos
would produce a current
(9.10)
which oscillates in lime as exp [ - i(wOI + WOS)tJ
v = Vt cos Wit
iwt
+ F·e- iwt
(9.18a)
where the· refers to the complex conjugate. For example, if we have a group of identical spins as in (9.7) acted on by a rotating magnetic field, we get Vet) = -,!tlll(J'Z:T cos wt - lyT sin wt)
(e eiwt
= -,"/till, [[Z'f := _
iWI
+2
)
+ ~IYT(eiwt _
,!i.ll, (ueiwt + l-e- iwt ) 2
T
T
so that, using the usual notation WI
e-
iwt )]
(9.1gb)
=..,1l I, (9.18<:)
435
Note that a rotating field applied in the opposite sense is obtained by treating
W
as negative. . Let the eigenstates of 'Ho be denoted by quantum numbers, J, k, 1 each of which stands for some sel of quantum numbers such as M, 0' of (9.7b). Then we
Substituting (9.18) for V/IJ in (9.24), we thus get
dl
exp(-iwj/t)....::l!..dI t
""
define
i
L:, ejk {exp [i(w - i
(9.19)
Wjk)tjFkl + exp [ - i(w + Wjk)t)FkI}
L: , {Fjk exp [i(w -
Wk/)t]
+Flkexp[ -i(w+wk/)l]}ekl
(9.25)
Then, denoting the interaction representation by a prime, we have
ee:O
= e -(i/A)11ot {!' (l)eCi/A)11o t
(9.20)
or
(9.21.) (9.21b) The density matrix then obeys the matrix equation d i dt {ljl = h ~ [ejk('Hoh:1 - ('HO)jk{!kEl
+ .!. L [ej' V,,(t) h ,
Vj,(t)e"J
(9.22)
If we had only the tenns involving 'Ho, (!jk would oscillate at exp(-iwjkt). We can eliminate them from the right by substituting (9.2Ia) into (9.22). We also utilize the fact that
The first two tenns on the right give the sum and difference of wand Wjh the third and fourth lenns on the right have the sum and difference frequencies of W with Wk/. This is the same situation we gOt for our example of the nonlinear currenT/voltage relationship (9.12) and (9.16) in which the product of V1(t) with V2(t) gave the sum and difference frequencies. Here what is happening is that we consider the elements of e' on the right hand (either e"k or ekl) to be the n..quantum coherence which we wish to pump by means of'v up to the (n+ 1)quantum coherence or down to the (n - I)-quantum coherence represented by ejf on the left hand side. The point is that since ell [the (n + 1)- or (n - I)-quantum coherence] is to be produced, its time derivative must be different from zero. Let us suppose, then, that at t "" 0 the matrix elements ejl and ek/' and thus ejl and e'kl' are zero, but that we have previously excited the mauix element ejko for a particular value of k, labeled ko. So, we have a nonzero ~jko' hence f!..krJ' Let us suppose funher that V(t) possesses a nonzero mauix element betkn states ko and 1. Then, we have
dt ""
de'· exp(-iWjft)
+exp[ -i(w+Wjko)t]Fkotl
(9.23)
.
' d (!'f
.
ht ""' L..(ejkexp(-IWjA:t)Vkl(t) I
.
Fkol "" 0
I
-
~ V,'k(t)eklexp(-iwk/t)]
"
(9.24)
This equation reminds us that the matrix elements of e' would be independent of time (Le. that dej/dt "" 0) if V were zero. Consequently, w~ expect that mauix elements of e' to change slowly compared to the frequenCies Wjl' Wjk. Wk/' Therefore, to a good approximation the left hand side of (9.24) oscillates at Wjf. This requires that the right hand side also oscillates at the same frequency if (dtl;/dt) is to be something other than zero.
,
Fiol i- 0
.
(9.27)
Then we would have exp (-iwjlt)
:!' "
d'
ielko F kol exp [ - i(w + Wj~'o)t]
(9.28)
Then the condilioll that F pump ejko up to Iljl is that F kol i- 0
and
(9.29.) (9.29b)
Wj -WI ""Wj -wko +w
436
(9.26)
Now, ordinarily, F is a raising operator and F* a lowering operator. Thus, recalling that ko and I will in general be specified in pan by the quantum number M, see (9.gb), we see that if Fkol i- 0, then Fiol "" 0, or vice versa. Let us assume
since the states, j, k, 1 are the eigenstates of 'Ho. We thus get exp(-tWjlt) d: =
iejko {exp [i(w - Wjko)t]Fkof
(9.29c) 437
pumpThis equation may be interpreted as saying that the beat between the wig) Wj ncy (freque ce ing term F (frequency w), and the n-quantum coheren [If ce. coheren tum l)-quan must add to give the frequency, Wj - WI, of the (n + ce.] tum coheren W is negative, the left hand side is an (n - I)-quan as written be also can (9.29c) n Equatio (9.29<1) .
Energy
Fig.9.J
....,
3
3
'·-1
2
Fig. 9.2
1--1
1-.)
W.&:o -WI::W
match The meaning of this is that to excite ejl from ejko (or ejko)' W must matrix a have must V and 1, to ko n transitio the of the resonance frequency which joins ko to 1. ce If we stan with a diagonal {J, we can first produce a one-quantum coheren two-a to up this pump can we that shows above with a 1((2. pulse. The discussion coherence, quantum coherence. That in turn can be pumped to a three-quantum transition one excites which rf a apply can etc. Note that we are here assuming we Thus, (9.25). of side hand right the on at a time since we kept only one term the out pick to weak ntly sufficie HI an if there are splittings, we are using l spectra between spacing the than individual lines. It must therefore be smaller strong A on. exciuu{ e selecliv cy lines. This mode of pumping is called frequen it in lhe HI covering all the splittings is called nonselective excitation. We treat next section. levels Let us follow our example in greater detail. We define three energy I, 2, 3 (Fig. 9.1) with the order of energies such that W3 >"'2 >wI
(9.30)
We assume that W3 -"'2
4-
"'2 -WI
(9.31)
not e32). and that we have previously excited 1!21 (hence e12), but not 1!23 (hence Fig.9.2. in spins of pair a for shown is system a such of ion An actual realizat Assuming lhey have a coupling of the form (9.32) rand 5, and are identical except for a chemical shift difference between spins ed in discuss we as cies, frequen t differen at occur the four allowed transitions Chaptet 7. Let us rewrite (9.24) by multiplying both sides by cxp(iw jlt). Then .I t del ./ i · (9.33) ~ :: h ~ [ejke+lI ·Ik/IVkl (t) - Vjk(t)e + "-'J1- I!I./] with the This is actualJy just what we would have gOllen if we had started (9.21)] ing [follow equation for I! in the interaction representation, with Vkl(t) :: Vkl(t)ci"-'k/t
etc. Then, 438
(9.34)
....
,
I•• )
2
...., h -' h ex""'---· E, E,,.. labeled 1, '1 , 3 wilh ener&ies E, " levels lergy <=eU3$ "-'I, l<.'2 ( . L_ . Threeel hFig. 9.1. ""3 -l<.'2 '# "'2 _ "'I' ""3 In til'" notallon of 9.19). Note that l<.'2 > l<.'2 > "-'I and we assume lhe lhree levels labeled Fig.9.2. The energy levels of a pair of spin-t nudf!i. We focus on 1, 2, 3
. it l
' d 1!31
--;It "" -he "-':)1
d
I
V32e21
and
~~I
.
I
""
-*eil<.'2 3tv23e; 1
(9.35)
F- the Usin~ (9.1&), we see that F involves the raising operator r+ + S+ and I + _) is two state and -) I is three lowenn g operator r- + 5-. Since stale ' we have
F32 ::: - ~I(- - I r +5-1 +-) =_wI (__ jrl+_ )
2
WI
(9.36.)
::--
2
+ ..... WI F32' -2(-I'· +5 1+-) ·0
(9.36b)
Likewise (9.360) Therefore
de;1 :: dt
,.w,2 ei(""31-,,-,)L
I e21
de21
) WI'( dt:: fTel l<.'23+"-' t e3t
(9.37)
The right hand side has an oscillation at W3-W 2-W
.
(9.38)
This causes lhe left hand sides to vanish unless 439
(9.39) If we satisfy (9.39) with an equality, then we gel
tP£'~lI
(WI)'2 £'31 - O ' dd'2£'21+(WI)2,=O t2 2 f'21 t
-;ur + "2
_
(9.40)
which we recognize as hannonic oscillator equations. Writing down the general solutions, then evaluating the constants of integration with the help of (9.37), we
gel
,
,
£'31(t)=e31(O)cos
,nd
(W"). , (0) Sin. (W") T +1£'21 T
e~!l(t) = e21 (0) COS (wt) + ie~l1(O) sin (w~
t)
(9.41)
lllUs, if we turn on Hi. e~lI will transfer its amplitude to l?31' which will then transfer it back 10 e~l later tInd so on. Note that if t is applied for a lime t w generating a 11" pulse (WI t w = 11") so that (9.42)
One can represent the basic equations (9.28) and (9.29), describing selective el(citation of (n ± 1)-quanlUm coherence from a state containing n-quantum. coherence graphically. as shown in Fig.9.3. We use a solid line between two energy levels to denote the matril( element of V joining those states. Dashed lines between states represent matril( elements of f' (or t/) between those states. Then, the requirement that to generate a "if" matril( element on the left of (9.28) demands a jkg and a kgl matril( element on the right simply means that the two dashed lines plus the solid line fonn a closed loop, as in Fig. 9.3a. Then one also automatically satisfies the energy conservation requirements (9.29c) or (9.29d). Note that although V is in resonance with the 2-3 transition, we have assumed it is not in resonance with the 1-4 transition since we did nOI connect those states with a solid line. In Fig.9.3b we show that if we tuned V(w) to be resonant at the 1-4 transition, it cannot pump f'21 into f'31 since the lines do nol fonn a closed figure. However, looking at Fig.9.3c we see that V"I will pump f'21 into ~24. the zero-quantum transition. Note also that these figures show (Fig. 9.3a) that V23 can pump l?JI> the twoquantum coherence. into l?21, a one-quantum coherence. Also (Fig.9.3c) VI4 can pump the zero-quantum coherence into the one-quantum coherence l?21. We began Sect. 9.1 with some general remarks about the density matril( of a system in thennal equilibrium, asking whether or not one could experimentally
(9.43)
l'31(tw) = iU21(O)
(al i4,(lw)=O
(since £'31(0)=01
(bl
.
Thus. the full coherence has gone from one-quanlum coherence to two-quantum coherence at time two These same equations show that if we start with pure two-quantum coherence [£131 (0) f- 0, e:u (0) 0). we can transfer it to one--quantum coherence [~21 (t w ) j. O. ~:H (t w ) = 0] by a 'If pulse (WI t w = 'If). It is interesting to note that, if one starts with pure one-quantum coherence and applies a 211" pulse (WI t w = 211") to transition 2-3 (flipping the I-spins)
2
~21(tw)=-e21(O)
(9.44)
Now,
~21 = (+
-1ll'1 + +)
(9.45)
gives the $-spin transverse magnetization. Thus. a 27r pulse applied to the Ispins reverses the sign of the S-spin transverse magnetization. This result is a manifestation of the spinor character of spin systems. It was employed by Stoll et al. [9.9, to] and Wolf! and Mehring [9.11) to demonstrate the spinor nature of a spin panicle. Mehring el al. [9.12J have utilized this fact in an ingenious version of spin echo double resonance for electron spin echoes when the electron is coupled to a proton giving a resolved hyperfine splitting of the proton.
i
i
440
\ 911.W)I
1
,
I
1
911 \
\
I
l
2
\
1 1
\
I"
,
1
'''\ I 1
\1
A
V41
(el l
1'2,(0)'; o. '3,(0) ~ OJ, •
it' W
\'{"'11
=
f'JI(tw)=O
Vn
2
("----!1\
" ,,,""" /< 1
\
,
VII
FIg. 9.3a-c The four energy levels of a pair of spin-t nudei. (a) The density matrix dements e~1 and 1<'31, which oscillate at l.o.I~1 and ""':II respectively, Rrc CQupled together by the lime-dependent perturbation V oscilla.ting at fre
detennine whether all eij (i • j) were zero for a system in thennal equilibrium. The fact that we can pump n up or down gives us hope. Thus if we want 10 inspect a panicular eij. we look for a sequence of transitions to pump which eventually enable us to pump the ij coherence (if it is nonzero) into a pair of levels kl whose coherence, t!kl, is directly observable. Alas. a simple example will suffice to show that the usual excitations will not do the job in all cases. Consider two identical spin nuclei (e.g. a pair of protons) with identical chemical shifts. Their energy levels are the singlet and triplet states (Appendix H). Designating Ihe total spin by F, we have
i
F= II + h
Fz = liz + 12z
(9.46)
=
Fz == I tz + I2z
(9.47)
has no matrix elements connecting the singlet 10 the triplet, the only nonvanishing Yij'S are entirely within the triplet. l11OS, if we wish to inspect a matrix element
(FMFlelF' MF') = (II1e100)
(9.48)
we could only pump it to stales such as (lOlt!loo) or (I - II£'Joo) in order to generate a closed set of lines. Figure 9.4 illustrates how we could use (Illeloo) to feed (IOleIOO). But (lOlelOO) is not observable either. nor is (I - lIeloo) which we could produce as in Fig.9.4b. Thus, we can only succeed in pumping our original unobservable matrix element (9.48) into other nonobservable matrix elements! In principle we could apply a spalially inhomogeneous ahemating field of strength HI at nucleus I and H2 at nucleus 2 so that we had a rotating frame Hamiltonian
Ibl
I---<
11, -1191001
11'-
II°'I---:<
lI
V
110191001
.
"':>:>-- 10.01
V
1111
--(1119 100 1
11,01
r--_ \
11019100)
"-
\
-.;1
10,0)
11,11---
Fig, 9.4. (a) An alternating potential joining the states J10) and lit) can couple (lotf?IOO) to (t1If?IOO). If we wanted to insp«tan element (tllf?IOO) whidl is itself not directly observable, we a~ uJ\$uccessfulsince we can only pump it into an element (tOI(lIOO), which is also unobservable. (b) If we produced (toteIOO), a furlher IIpplieation of \' tuned to the JI- t) -110) trllnsition will produce (t -lleIOO) which is still nOl observable 442
=
--yhhOFz + hahrh: _ -yh( HI; H2 )Fz
--yl{ HI; H2 )(lb - hz)
(9.49)
The tenn Itz - hz now connects singlet to triplet states, as can be seen by evaluating matrix elements between Ill) and Joo) using the spin-up and spin· down functions it and fJ fO express the IFM F) Slates:
III) = ,,(1),,(2) .nd
=
For the two spins, we have F = 0, M F 0 for the singlet, F I, M F = I, 0, -I for the triplet. We now demonstrate a type of matrix element of t! which we cannot probe. Since
101
1ieff = --yllhoFz +haIlzl2z --yh(HIIlz +H2I 2z)
(9.50.)
I
1(0) = j2[,,(I)P(2) - P(I),,(2)]
(9.50b)
We find that
1
(II Ill. - 12' 1(0) = - j2
(9.5Oc)
Thus, if we had initially a nonzero malrix clement (10IU
to produce
en,
(9.51) then pump the 23 transition using
W=W32
to produce e13. This would require shifling the frequency W from W21
(9.52) 10 W3to
443
something of a nuisance 10 do. In practice, one usually works with large HI's which can flip spins corresponding to many multiplet lines. Such pulses are called nonselective. They cover all the lines within an angular frequency width Aw="IHI
(9.53)
of the oscillator frequency w. We turn now 10 see how they work.
9.3 Nonselective Excitation
hence that there were no groups of three protons within a given molecule. They thereby showed that the species CCI-I3 was not present. 9.3.2 Generating MuUiple Quantum Coherence To generate multiple quantum coherence in the absence of detailed spectral information we utilize an approach which is referred to as nonselective excitation. There are many variants. We stan with the simplest. We will first describe without explanation the pulses we apply first to produce, then to detect the generation of multiple quantum coherence. Then we will explain how the sequence works.
9.3.1 The Need for Nonsclooive Excitation As we have seen. using frequency-seleclive pulses we can pump an n-quanlum coherence either up or down. Thus. slaning with the usual one-quantum c0herence following a 7fo. pulse. we can proceed in principle to produce a pure two-quantum coherence. then conven that 10 a pure three-quantum coherence. and so on. To do this, we would need (0 know the energy levels and the cor· responding wave functions. Having produced some desired. level of coherence. we could later inspect how it evolved in time by later reversing the process and pumping the n-quantum coherence to n - I. retuning the pump frequency to pump that to n - 2. etc.• evenlually gelling back to one-quantum (observable) coherence. If we studied how the observed coherence varied with the length of time the system was in the n-quantum state, we could learn about Ihe evolution of the system in the n-quantum stale. However, this approach implies we already know the energy levels of the system, and know therefore what spectral lines are connected with what pair of states. This situation would be satisfactory for some purposes - for example. if we wanted to verify that the two-quantum matrix elements of e of a pair of spins vanishes when Ihe system is in thennal equilibrium. However, frequenlly one is attempting to utilize multiple quantum coherence to understand some unknown system for which the spectml lines may nm yet be assigned. One then wishes some general approach to exciting mUltiple quantum coherences. One would then leI them evolve. and lastly inspect what happened during the evolution, to try to learn about the system. For example, one might wish to know whether there are groups of two spins but not groups of three, as in the experiments of Wallg et al. [9.13]. who studied acelylene (C2H2) adsorbed 011 the surface of Pc metal and wished to detennine whether or not any CCH3 spccics wcre fonned. They found they could genemte both two· and three-quantum coherence. By comparing the intensity of two-quantum coherence wilh that of three-quanlum coherence. Ihey showed that groups of three or more prowns were rare. By studying how the relative intensities of two- versus thrcequantum coherence depended on the concentration of acetylene. they showed that the three-quantum coherence arose from coupling between different molecules,
444
.
preparation
evolution
detection
Fig.9.5. A three.pulse sequence, sho"'inS the prc,r'rlltion, e\'olution, ods, lind defining the times 0- ,0+, T - ,T+. r;, ' I
111m
detection peri-
The simplest approach to generating multiple quantum coherence is illustrated in Fig. 9.5. A strong 7f12 pulse is applied to a system in thermal equilibrium. "Strong" means "IH} exceeds the spectral width• ..1w, of the spectrum. That is. HI is larger than the spectral spliuings arising from chemical shifts or spin-spin couplings. The system develops freely for a time T. This pericx1 is called the preparation period. During this period. as we shall see, spin-spin coherences develop so that when a second rro. pulse is applied at the end of T, it convens the spin-spin coherence to a variety of multiple quantum coherences. These coherences then develop freely for a time tt. called the development time. At the end of that time, one wishes to inspect what has happened. This is done by applying a rro. pulse which transforms the multiple quantum coherence back to one-quantum coherence for detection. The signal is recorded as a function of Ihe time t2 after the detection pulse. Frequently t, is varied as a parameter to enable one to identify the contributions of the various multiple quantum orders. Thus, the signal, S. is a function of both t I and l2. It will also depend on the preparation time, T. giving us in general a complex signal, S(T. tl, t2)' To explain how the above pulse scqucncc works, we considcr a model system of two identical spin ~ nuclei, spins [ and S, possessing identical chemical shifts. We suppose that the rf pulses have a frequency wand write the Hamiltonian in the coordinate system rolating at w. In that frame, in the absence of the pulses, the Hamiltonian is thcn (9.54,) 445
n == "(Ho -
M
(9.54b)
w
representing the possibility of being somewhat off resonance either from the existence of chemical shifts. or because the experimenter has deliberately chosen
coupling. See Appendix H. For more than
twO
spins. the form of the interaction
term ([lS:) is only an approximation even if the coupling is dipolar. However, though Ihis fonn of coupling gives only approximate results in thai case, it makes possible simple explicit calculalions. enabling us to get results whose physical
significance we can explore. We utilize the spin-operalor method (Sect. 7.26) to follow the density matrix in time. DenOling times 0-, 0+. r-, r+, etc. as indicated in Fig. 9.5 to indicate times just before and just after the pulses at t = 0, T, etc., we take the density matrix at t = 0- to be e(O-) = (l, + S,)
(9.55)
omitting the constant of proportionality. Then the development of e is given in Table 9.1 up to the time just after the second pulse. We utilize the shorthand notation cos (0,/2)
==
C;
sin (nT)
Sh '
==
etc.
(9.56)
It is useful to express the spin components I z , Ill' etc. in terms of raising and lowering operators such as
I
1
I. = -(P" + r) , I, = 2', (P" - n . 2 We have collected the results (or products of twO spin components in Table 9.2. Utilizing these results we get the last column of Table 9.1, which expresses e{,+) in terms of raising and lowering operators. Utilizing them, we see that line 1 of the table conesponds to elements (m,mslelm'/ms) for which T.ble 9.1. The lime dependence of (/ durin, Ihe preparalion period Column no.:
Drivin,lcrms: Column labc:I~: (/(0-)
2) X(1I'/2) (/(0')
/,+ 5,
4 aT/,S,
5
,
,
m, +mS
(I,+S,)ShC~
(9.57)
(9.58)
hence are one-quantum mauix elements. Line 3 has products of two raising operalOrs or two lowering operators, hence corresponds to M'=M±2
,
(9.59)
which are two-quantum matrix elements. Lastly, line 4 has I z or Sz multiplied by either a raising or a lowering operator, hence has
M'=M±1
(9.60)
again a one-quantum coherence. Utilizing Table 9.2 we see that at t = 0-, the density mauix consists entirely of zero-quantum coherence. It is converted to one-quantum coherence by the first X(1rI2) pulse. At t = T-, just before the second X(1r!2) pulse, we note £:I(T-) still consists entirely of one-quantum coherence. However, there are now (our terms (lines I through 4) as opposed 10 the single term at t = 0+. Half of these terms arise simply because we are off resonance. That is, they come from the effect of the off·resonance term n(Iz + Sz). The other half arise from the effect of the spin-spin interaction term alzSz . For lines I and 2, the elements of £:I are proportional to cos (o.T(2). For shon T, those terms survive even if a goes to zero. Thus, they would be present even if I and S were nOt coupled together. In the limit of uncoupled spins, line I corresponds to magnetization which is flipped by the first pulse into the z-y plane, and then is flipped back onto the z-axis by the second pulse. Note that as , goes to zero, the combined effect of the two 1r(2 pulses is to flip the magnetization onto the negative z-direction. They are thus equivalent to a single 1r pulse.
•
(/,+S,)ChC~
' m,'+ mS
M'=M±1
,
X(III2)
D.(/,+S,)
"1' -=
HI
are the same (M' = Ivl), hence are zero-quantum matrix elements. For line 2, the terms have
to be off resonance. Recall that this form of coupling would give exact results for a pair of spins with dipolar coupling, bUI would not describe twO spins with a pseudo-exchange
==
"
-(l,+S,)ChC~
Line
-(/,+S,)ChC~
..
~(r +s'+r +s-)ShC~ 2
__ 2..(1' S' - r S-)Ch2S' 4i
•
2
,
(/, 5, + I, 5, )Sh 25:
446
447
T.b~ 9.2. ProdllCl$ of 5pln oomponcnu cKprcs5Cd in lcrms of r.lslnl .nd lowcr;nll opcr.tors
I.S~ .. ~/,(5' +5-) 2
I,S,
_~/,{S' -5-)
5,1,
_~5,(r +r}
...
S,I,
_~s,(r -r)
I 5
"~{/'s'+rs-+rs-+rs')
IS
--~(I"s'-rs'-rs-+rs')
.
" IS .2.(r5'-rS--I·S-+r5') >'
Line 2 for a = 0 corresponds 10 two uncoupled spins which are flipped inlO the x-y plane by the first pulse and left Ihere by me second pulse (because mey point parallel to HI of the second pulse). On the other hand, Jines 3 and 4 are proportional to sin(aT/2). Therefore, if one swilches off Ihe spin-spin coupling (a _ 0), these tenns vanish. They do nOI therefore arise for uncoupled spins. In order to observe these terms, one mUSI choose a time T such Ihal sin (aTn.) is appreciable, hence T'S for which aT/2 ::::: 1rfl, 31fn., elc.
(9.61) 3 If a represenlS dipolar coupling as in solids, a ::::: I/r where r is the distance between spins I and S. Thus, by selecting T, one can pick out the values.of.a or of Ihe distances T which will contribute. For example, the proton pairs 10 a CH2 group have a characteristic H-H distance. The existence of such a pair will give rise to a characteristic splitting of the proton NMR line into a pair of lines whose separation is proportional to the coupling. This splitting, first observed by Pake [9.14] is often referred to as a Pake doub/~t. Our a.naly~is shows that the existence of the coupling, hence of the doublet, Will also gIVe nse to the possibilily of generating double-quantum coherence. The pulse spa~ing r to produce the maximum coherence depends on Ihe strenglh of the couphng, a, hence on the magnitude of the splitting. Note that one can distinguish coupling between protons in the same molecule from coupling of protons in diffe~nt molecules either by setting T to correspond to the desired distance, or by.studYlOg the effect on the Iwo-quantum signal strength of diluting the molecules 10 a nonproton-containing matrix, as in the experiments of Wang et al. Extensiv.e SlUdies of this sort have been carried out by BQllnl et al. [9.15], demonstrating large clusters of protons in a given molecule. 44.
We note that while line 3 gives double-quantum coherence, line 4 gives single-quantum coherence. The distinction between these two lines arises solely from the precession effect of being off resonance. The lines are identical except that the tenn cos nT of line 3 is replaced by sin nT for line 4. If one were exaclly on resonance (n = 0), line 4 would be zero. However, one could men excite this single-quantum tenn by making the second pulse be Y(1f/2) instead of X(7ffl). In the process, me double-quantum coherence would vanish. The point is mal if one is exactly at resonance, the phase of the second rf pulse will determine whemer one gets single· or double-quantum coherence. We discuss these points more in lhe next section. 9.3.3 Evolution, Mixing, nnd Detection of Multiple Quantum Coherence We have now seen how Ihe first two pulses produce zero-, one- and two-quantum coherences. We now wish 10 see how these develop in time. Moreover, since neither zero- nor double-quantum coherences are direclly observable, we want 10 see how me third pulse converts them to something we can measure. The conversion is conventionally called "mixing". Since we observe nuclei by singlequantum coherence. we therefore wish to see how the third pulse converts zeroand double-quantum coherence to single-quantum coherence. We merefore conlinue our study of how fJ develops in time under the combined effect of me Hamiltonian of (9.54a) and a mird (mixing) rf pulse at time tl after the second pulse. Using the spin-operator methods, we construct Table 9.3 to describe the evolution and Table 9.4 to describe the detection periods. From me tables we can follow the effecl of the differenr interactions (spin-spin and Zeeman) on me time development of the density matrix. We now discuss some important features the tables reveal. First, we back up a bil 10 express the lime development of fJ. During Ihe time interval between pulses, the Hamiltonian (9.54a) is static, hence the lime development of the density matrix obeys an equation similar to (5.85)
("1,(1)1"') = ("Ie(O)I,i)e,p(i(w,,, -
w,,))
(9.62)
where (9.63)
W mrms
= -n(m, + mS) + wnJmS = -{}M + am/ms
(9.64)
where M=mJ +ms
(9.65)
gives the total z-componenr of spin angular momentum. Utilizing (9.62) we see then that the time dependence of fJ between pulses is given by
449
Tablt:
'.J. Time devdopmml of g durilll tM r:¥OIulion and mWlII period
Number of quanta in column 1
o
,
Column no.; I Column la~l; g(r') Operalor; Line number
•
)
-(/,+S,)ChC;
(I.+S.)SoC~
•,
•
-(I, S,+I,S,)C:Sb 2S~
-(I, S,+ I,S,)Sb2S~
'Q
'Q
450
'Q OQ
-{I, S,+ I,S,)C~ 2S:S o C;
(/~ S, + I, S,)Sb2S;SbC;
'Q
-(I. S,+ I,S.)Sh2S:S o C;
'Q
-(I, S,+I, S~)C!QCO 2S;
'Q
+(1, S,- I.S.)S;QC O2S~
O,2Q
- (/, S,+ I,S,)C~C;So2S~
(I, S, + I, S,)ChC:So2S~
+ (/,S~+ l~s,.)ShC:So2S~
'Q 'Q
+ (/:S~+I. S:)2S:Sg 2S; {S~+I.)Chs:sgs:
- {S~+ 1.)S;SoS;
(S~+I.)chS;SbS;
-(S,+I,.)ShS;Sos;
Line 00.
,, ,• • , • "" .."" """ "" "" "
Q
..
Table 9.4 (conlinued)
during llle delecticn period
,
Column no.: Column la~l: Q(r+f,') Operator:
+(I,+S,)CoC;
2
•
l
(I,+S,)ChC;SbC;
-(S,+ 1,)ShS;sos:
'Q OQ
, Q(r+fl+'V
"'I{/,S,)
+ (/,+ S,)~CoC; -{I. S, + I, S.)2S:CoC~
+(I~+ S.)SbC;CoC~
{/.+ S~)C:ChC;SbC~ (I,S, + S,I,l2S:ChC;SgC:
7
'Q
'Q
(/,S,+I, S,)Cb2S;SbC;
-{/,S~+/~ S,)ShC;So2S~
)
'Q
+(I,.+S,.lCoC~
Num~r of quanta represented by column 1
- (I.+S,)S~C;SbC;
+(/, S,.-I. S,)2ChShCb 2S;
durilll evolution
'Q
•
- (Shlll c)., 2S;
Table 9.4. Time developmenl of
OQ
, g(r+ft)
(I~+S~)ChC;soc:
-(/~ S,.+ I,S~)I{C~ll
- (/.S,+ I,S.)C o 2S~
"
o>
(/~+S~)C~C;ShC;
(I, S, + I, S,)2S;SoC;
• • "
co~rC11o«
-(/,+S,)ChC; -(/,+S,)S~C;SbC;
7
Order qUilnlum
• X(II/2)
{I. + S.)C;ShC;
)
,
,
,,(HI,)
01,(/,+1,)
-{/,+S,)CoC;
,
Table 9.3 (ronlinued)
-{/,S,+ I, S,)C~ 2S;SDC~
- (I, S,+ I,S,)c:ch2S:S o C;
+(1: S,+/~S:)2S:C:1I2S:SoC; - (I~ S,+ I, S~)ClDCO 2S~
-{/:S,+I,.S:lSb2S;Ch25;SoC; .. -(S,+I,.lsbs;Chs:SbC;
-(/.S, + I, S~)C;ClQCo2S;
-(I, (I,S,+ I, S,)ChC;So 2S;
(I;S~ + I. S:)cb2S;Ch2S;soc: _ (S. + 1.)CbS;Chs;5 0C:
S: + I: S,l2S;C;QCo2S~
(I,S,+ I,S,lC;chc~So2S;
- (I. S:+ I:S~)5b2S;ClD25;
.. - (1,+ S.)SbS;C1QChchS~
-(I. S:+ I:S.)Cb2S;ChC;So2S~.. -(l~+S.)cbs;chc:sos;
- (I.S: + I:S.)2S:ChSg2S; (S~ + 1~)Chs:sDS:
{S~+I.)C;Chs;sos:
(S,I,+ I,S,l2S;chs:sos:
451
exp(i[n(M - M ' ) + a(m~mS - mrms)]t) = exp (W(M - M')t)exp (ia(m~mS - mrms)t) (9.66) where the first factor arises from the Zeeman offset. and the second from the spinspin coupling. Recalling thai n = 1M - Mil gives the order, n. of the coherence we see immediately that the time dependence of the frequency offset uniquely labels the order of coherence of the element_ This fact is evident in Table 9.3. The time dependence of the Zeeman offset is given by the terms involving and t 1. They are to be found in column 5. and are re-expressed in column 7. Since M - M' = 0 for zero-quantum terms. (M - M')ntl = O. hence the Zeeman offset term is merely a conslant independent of or 1I. This prediction agrees with the table where the zero-quantum terms (line I) have no terms depending on ntl' The one-quantum terms (lines 2. 3. 4. 5. 8, 9,10. II) all involve either Ch or Sh. The two.-quanlUm terms (lines 6 and 7) involve (Ch)2 - (Sh)2 = C~o or 2ChSh = S~n' hence are at frequency 2n. A second important fealUre of the table can be seen by comparing column 3 with column I. showing the effect of the spin·spin coupling (column 2) on ~. Examination of line 6, the matrix elements corresponding to two-quantum coherence, shows that the spin·spin coupling has no effect on ~. To get this result using the operator method, one gets eig1l1 terms. They can, however, be added together. Using relationships such as sin 2 +cos 2 = 1 and I~[11 = iI~/2, they can be reduced to a single final term identical to the starting term. One naturally wonders if there is some deeper significance - indeed there is. The effect of the spin-spin coupling is given by the second factor of (9.66)
n
n
exp(ia(m~ms - mrms)l)
•
(9.67)
which. for the two-quantum term, has mr
= ms = i
and
m~
=
ms = -i
(9.68)
or vice versa. But
am
m=a(-!l (-!l
(9.69)
Therefore, the spin·spin factor becomes and we get no difference on line 6 between columns I and 3. If one has a system with more than two spins. a similar theorem is true. Consider a general system with N spins of and a Hamiltonian in the rOlating fmme 1{ = - L hni1zi + L(AijIi' Ii + Dij[zi1zi) (9.70) ij
!.
corresponding to various chemical shifts and a generalized secular portion of the spin-spin coupling. The total angular momentum operatOr (9.71) with eigenvalues MF commutes with 1{. 452
Thus, we know that the state in which all the spins point up (with MF == N/2) and the state in which all spins point down (M = - N /2) are both eigenSlates. Using a notation Iml.m2.m3• ...• mN) for the m-values of spins 1, 2. 3, etc,. we know that the state IN/2) in which all the spins point up (i.e. with M = N!2) is given in terms of individual spin m values by (9.720) and the state with all spins point down is
1- N/2) =I-!. -!..... -!)
.
(9.72b)
It is easy to show by explicit evaluation using (9.70) that these two states have Zeeman energies which are equal in magnitude but opposite in sign, whereas the dipolar energies are equal. Therefore. the energy difference between the states. which gives the time dependence of the N -quantum coherence. is independent of the strength of the dipolar coupling. We turn now to the detection method period described by Tables 9.3 and 9.4. This requires converting the various orders of quantum coherence back to the observable one-quantum coherence, the step called mixing. Since a 1r(2 pulse initially convened one-quantum to n-quantum coherence, it is natural to ask if ~ would reverse the process. Indeed, in Table 9.3 we analyze the effect of an X(1f!2) pulse, which is just the inverse of the original X(7f/2) pulse. The result, in column 7 of Table 9.3. is labeled in column 8 according to the orders of the resulting quantum coherence. For example. line 3 represents zero-order quantum coherence. II would be undetectable. In Table 9.4 we keep only the coherences which are single quantum during the detection interval, but label the rows by the order of the quantum coherence during the evolution period. In column 3 there are tenns such as (l~S~ + I~Sz) which eventually give no signal because when we form Tr {g(J+ + S+)} the [z and Sz cause the trace to vanish. We do not therefore bother to follow these terms past column 3. In column 5. we include the effect of the Zeeman offset. 11le signal is observed as a function of 12, and is given in tenns of F== 1+ 5 as (9.73) Utilizing (9.73), we can combine various lines of column 5 in Table 9.4 to get the contributions of the zero-, one-, and two-quantum coherences to (F+). Denoting
by (9.74) the contribution of the nth order quantum coherence, we get for the zero-quantum coherence
(P+}o = ~[e-iCn+(J/2)/1 + e- i (il-(J/2)I1) x [eiCn+(J/2)T + eiCil-(J/2)T +e- i (O+(J/2)T +e- i(O-(J/2)T)
(9.75) 453
For the one-quantum coherence
{F+>' = 2e- iS7t2 cos [a(t} - r - t2)/2] cos (ntd sin (f}r)
the conversion of the initial z magnetization to one-quantum coherence for the preparation period, the fact that zero-, one-. and two-quantum coherences have been brought into existence for time interval tl' and the conversion back to single-quantum coherence for observation. We have seen that the pulse sequence X(-rr/2) ... r ... XC1r/2) will produce zenr, one-, and two-quantum coherence. As we remarked earlier. if we study column 10 of Table 9.1, we note that the zero- and two-quantum coherence terms (lines I and 3) are proportional 10 cos f}r, whereas the one-quantum terms (lines Therefore. if = 0 (exact resonance) we 2 and 4) are proportional to sin would not produce one-quantum coherence. The question, how, when we are tuned exactly to resonance. can we produce one-quantum coherence with a pair of spins. turns into the question. how do we produce odd-order quantum coherence for an arbitrary number of spins. when tuned exactly to resonance? We turn now to a simple approach. The disappearance of the one-quantum coherence for a pair of spins can be traced back to column 5 of Table 9.1 where we see the effect of the spinspin coupling alone [i.e. before the resonance offset term f}r(1z + Sz) acts]. It produces (ly + Sy)C; on line I and -(lz:Sz + IzSz:)C:; on line 3. These tenos, if acted on by X(1r/2), produce (!z + Sz)C; (a zero-quantum coherence) and -(!z:Sy + I y S z )2S:; (two-quantum coherence). To get the one-quantum terms we allowed f}T(lz + Sz) 10 turn I z into -Iy • I y into I z , then we applied an X(1r!2) pulse. Therefore, if we have f} = 0, so that I z and I y are not rotated by 1r/2, we should get one-quantum terms by rotating the pulse from being an X(1r!2) to being a Y(1r/2) pulse. Indeed, we see that a y(7I'"!2) pulse would leave (Iy + Sy)C; alone as a one-quantum coherence. but tum -(IzSz + I z S z )2S; into (IzSz + I z S z )2S:;, a one-quantum term. In general, of course, f} =f O. Then the points we have just discussed show that production of a particular order of quantum coherence depends on the amount one is off resonance. That is frequently a disadvantageous situation. It can be remedied by making f} effectively zero by means of a spin echo, then choosing the phase of the second pulse to get the desired order of coherence. Thus. we can use a pulse sequence
(9.76)
which, in complex fonn, gives
(F+)I = ~[e-i(n-aj2)12(ei(!Haj2)tl +e-iC l?-aj2)1 1)
x (eiCS7+aj2)T +e- i(S7-a/2)T) + e-i(J?+a/2)t2(ei(l?-aj2)t1 + e- i(l?+aj2)t 1 ) x (e i(S7+a/2)T +e-iCS7-a/2)T)]
nr.
For two-quantum coherence we get
(F+)2 = _2ie- int2 sin (at2/2) cOS(2f}tl) cos (f}r) sin (ar(2) = ~ [e-icn+a/2)t2 _ e- i(l?-a j 2)t2]
8
x (e2int1 + e-2int1 )(ei(S7+a/2)T _ eicn-aj2)T +e-iCS7-a/2)T _ e-icn+a/2)T)
(9.77)
Referring back to (9.62), we recognize thac the elements of the density matrix will have time variations
e+ i(l?±a/2)t
(9.78.)
for the states M - M' = +1, will have e- i(l?±aj2)t
(9.78b)
for M - M/ = -1, will have e 2int
(9.78<:)
forM-M/=2,and e~2int
(9.78<1)
for M - M/ = -2. We know from the Hermitian property of f! that whenever we have generated an M - M/ = p matrix element, we have also generated an M - M' = -p element. We note next that all three {F+)fl'S have the factors exp [ - i(f} + a!2)t2] and exp [ - i(f} - a(2)t2]. This corresponds to the fact that during t2 the spins are precessing in a given sense. We can identify the role of the preparation interval by looking at the rdependent terms. Note that these all involve single-quantum terms since their time dependence involves the frequencies (f} ± 012). We note that the time dependence tl tags the evolution period, and the order of the coherence is immediately obvious from the f} dependence which is exp(±Oif}tl) for zero order, exp(± Iiiltd for single order, and exp (± 2intl) for double-order quantum coherence. Thus, if one studies how (p+) varies with tl. one can isolate the zero-, one-. and two-quantum coherences. The expressions for (F+)p show clearly
n
X(,/2) . .. T/2 ... X(,) ... T/2 . .. [Y('/2) or X(,/2)] to produce the multiple-quantum coherence. One can also then eliminate the offset effects during the detection period by introducing a 1r pulse Cfor example at t2 = T12) to give an echo (at t2 = T).
9.3.4 Three or More Spins We have seen that a pulse sequence X('rr/2) ... r ... X(1r/2) will produce zero-. one-, and two-quantum coherence from a pair of spins if T is sufficiently long. What happens if there are a larger number of coupled spins? Will this sequence produce higher orders, or must one add more pulses? We shall show that the sequence still works.
454
455
1
i
We approach this problem two ways, first by considering three spin nuclei with a coupling of the form haikljrlh. which we can solve exactly, then we will look at the general case which we will treat by an approximation method. Let us stan then with a Hamiltonian for three coupted spins which generalizes that of (9.54a):
1t:::: -hillzT + h(aI2I%1Ir 2 + a23Ir2Ir3 + a3lIr3Irl) where IaT :::: Icrl + I cr2 + In 3 a:::: :to y, Z • (9.79) (Note, we shall assume below that the coupling between any pair, such as al2. can also be wOllen as a2l.) As can be seen from the case of two spins, it is the spin-spin coupling terms which lead to multiple quantum coherence. Therefore we need to focus on the second term of (9.79). Then, as with (9.55), we take e(O-), just before the first pulse, as (9.80)
,,(0-): I'T so that at t :::: 0+, just after the first X(.,./2) pulse,
(9.81)
e(O+) :::: IlIT
M::::M'±1
-*1tu t) IzT exp (*1l u t)
where
Thus all are one-quantum operators. The frequency off-set term in 1i, -hillrT' will cause a precession of the x- and y-eomponents of spin, as can be seen from column 7 of Table 9.1. It will leave the Irk (k:::: 2,3) terms alone. but will conven 1%1 or Iyl inlo linear combinations of IZI and II/I so that we should replace Izl by
(9.82)
exp (-io 12/r I I r 20 exp (-ia23 Ir2Iz3t) exp (-ia31 Ir3/%1 t)/1/1
(9.84)
In (9.84) the exponentials involving I r2Iz3 commute with Illb hence annihilate one another so that (9.34) reduces to exp (-iaI2/r 1I r20 exp (-i03 I Iz3Ir 1t)Iyl X exp (i031 Ir3Irl 0 exp (i012Iz I Iz2t)
(9.85)
We can now apply either of the (I,3) operators to get a result similar to column 5 of Table 9.1. Then we apply the (1,2) operators 10 that result. TIle final result is
(*1i t) u
: cos (a31T/2) cos (a12T!2)IlIl -
cos (a3IT/2)2 sin (aI2T/2)I%l/z2
- 2 sin (a31T/2) cos (a12T/2)I%l/r3 - 2 sin (031T/2)2 sin (012T/2)Iyllr2Ir3 456
(9.88)
at + 1%1
sin at
(9.89)
(9.83)
Since IlIT :: IlIl + 11/2 + 11/3. and since all three spin-spin terms commute with each other. we have to compute quantities such as
x exp (ia31 Ir3Id)exp (i023Ir2Ir3t) exp (iaI2I%1/r20
m
and replace I lIl by II/I cos
?t,,:IE h(0l2IzlIz2 +023Iz2Ir3 +03IIz3Izl)
ex p ( -*1iut)Izl eXP
(9.87)
1;1:) cos ilt - I yl sin
Then we wish to calculate ex p (
We have written (9.86) in a manner intended to help indicate how, using the spin operator formalism. each term arises. We note first of all, that only spin I has either x- or y-components. Secondly, we note that only coupling constants 0t2 and 031 involving spin I occur. This is because terms such as 023Ir2Ir3 commute with Iyl' Thus. if we had four spins, we would get from I yl terms involving 012, 013. and al4 but no terms involving 023, 02~, a3-l. We note that we can easily find the contributions from I y 2 and I y 3 from (9.86) by cyclic permutation of the indices. Expressing Izl and Iyl in terms of raising and lowering operators shows us that all terms of (9.86) connect states M and M' which satisfy
giving us for the contribution of I zl to e(t) cos (031 T(2) cos (0I2T12)[ cos (GOlyl + sin (ilOI%d +cos (a31 T/2)2 sin (OI2T(2)[ cos (at)I zl I r2 - sin (ilt)III1 I z 2]
e(0)l:.. 1
:
+2 sin (03 I T/2) cos (012T/2)[COS (ilt)I>:\I z2 - sin (nt)Iyl l z 3] -2 sin (031 T/2)2 sin (012T/2)[ cos (ilt)II/I Ir2Ir3 + sin (nt)Iz1Ir2Ir31 + tennsobtained by permuting the indices 1,2,3
(9.90)
This Hamiltonian is now acted on by the X(rr/2) pulse. In order to see what spin coherences are thereby produced, we need only examine what happens to the spin factors. TIle result is a table of operators before and after, and the order of quantum coherences which can arise from each term. Denoting the values
M-M'::p for nonvanishing (M O'lulM' 0'1), we get Table 9.5. The various possible orders of quantum coherence can be deduced by expressing the spin operators as linear or (k '" 1,2,3), then multiplying them to get terms combinations of I r /,:, such as for line 4
It, 1';
(9.86) 457
(9.93)
Tlblt 9.S. Effect of X{1l/2) pulst on operators Linc
Operator tltfOTe X(lll2) pulsc
Operator after X(>rI2) pulsc
,
I"
-I"
l
• ,
Ordcr!; of quantum collcrencc, p
0
,
I"
Ix, .. !(I,'+/,)
1.,1,2
,x,/~2" 2.. (I,' + I n(l; -/,-)
"
-1"/~l· -~/,,(1" -/,-)
1~1 "l
lx' I,l •
lx' 1<)
2.. .;
(I" +
/,-)(11' -11>
1,,/<)
,
-I"/~I" -~/,,(/1'-ln
I" /,//'1
-I"
8
I", 1,1/'1
6
which, using (9.92), is produced more quickly. 1J1erefore, if all spins in the 3 system are close together, we can produce the onc-, two-, and three-quantum coherences when the strong coupling Qjk'S satisfy
1~2 I~I· -~ I" (l i
-1 2- )(I{ -Ii)
Ix,IY1IY1·!(/,' +1'-)(/2' -/ 2 )(/1' -Ill
"
ajkT/2
:1:2,0
';;t
(9.94)
1r/2
If, however, two spins are close, the third far away, two of the three coupling constants are small, so that while two-quantum coherence can be produced quickly, three-quantum coherence is produced slowly. We can also analyze the two-pulse method for nuclei whose spin is not restricted to I = ~,and for arbitrary fonns of spin-spin coupling, such as dipolar coupling instead of coupling such as aIzSz , by using a power series expansion. Suppose then we have a Hamiltonian in the rotating frome
"
:1:2,0
"
:1:2,0 ±l, ± 1
1-l = -hfllzT +
8
L
1-ljk == 1-l: + 1-l88
(9.95.)
j>k 1 + It 4i II 2
p= +2
0'
.!-I+ r 4i 1 2
p=o
0'
1 4i II Ii
p= -2
where the fl represents being off resonance and 1-l88 and 1-ljk'S are the secular part of the spin-spin couplings. That is
[1-ljk' IzTJ =0
,
(9.95b)
[1-l 88 ,I:Tl =0
For example, for dipolar coupling
1-l j k = B j k(3I:jlzk - Ij' h)
(9.91)
(9.96.)
= B j k(2Iz jlzk - IZjIzk ~ IyjIyd
Once one has made such a table, one can work out the result in one's head since an Izk or an Iyk gives ± I, so a product of two such operators gives p values of 1 + I = 2, 1 - 1 = 0, -I + 1 = 0, -I - I = -2. We see that the three spins give us p values ranging from +3 to -3, as we proved earlier. Now, however, we can see how long T must be to produce Ihe coherence. Note that if a12 ~ GI3, that the three-quantum coherence will develop in a time when sin (aI2T/2) ~ 1 or UI2'/2 ~ 7r/2. If al3 «: al2, then we must wait until a13T/2 ~ 1r/2. In the process, as time develops, sin (aI2T/2) will have been through several oscillations. Thus, we must wait for the weakest coupling. Suppose, however, we had three spins in a line so that spins 1 and 3, on opposite ends of the chain, are weakly coupled. Then we expect (9.92) Then our expression (9.90) would imply GI3 would control how long it would take to generote a three-quantum coherence. Thai conclusion would, however, be wrong since, had we analyzed the time dependence of I z 2 instead of I zl ' we would have gouen a three-quantum coherence dependent on 458
(9.96b)
As in (9.7b), we then have quantum numbers Al and
IzTIMo:) = MIMo:)
and
0:
where
1-l ss IMo:) = E(> IMo:)
(9.97)
Then, the time development up to just before the second X(1r:I2) pulse is given
by {?(r-) = exp (
-*
1-lr) X(1rI2)I:TX(rr/2) exp
(*
=ex p ( -k1-lT) IyT exp (*1-lr) = exp (ifhIz) exp ( -*1-l88 T)IYT exp
1-lr) (9.98)
(*
1-l88 T) exp (-iflr I:)
The spin-spin exponential can be expanded using the theorem (see for example [9.16]) eABe- A = 1 +[A,BJ+~[A,[A,BlJ+
(9.99)
to give 459
e(r-) = e+
+
iI7r1
• { fYT - *r['Hss, IyT)
(.!.)2 ..!..-[1i,u, [1i [yT]J + ... }e tl 2! 88
•
(9.100)
-inrI,
Consider now the first commutator
[H.., IyTI =
L ,
(9.101)
[Hjk,!,,]
i>k
Equation (9.107) is, however, the condition for one-quantum coherence. In a similar way, one can show that every tenn in the series (9.101) generates onequantum coherence, a fact to which we will return. Returning to the two-spin Cllse, (9.100 and 103), we can now apply the Zeeman operators exp(inTI.:T) to get for the density matrix at time Te(T-) "" [yT cos DT + [zT sin DT -
~ fl I2[Izl [.:2 cos DT
- [yI I t 2 sin Dr+I.: I [z2 cos Dr-IdIY2 sin Dr]
If 1 is different from both j and k. the commutator vanishes. Thus, we have typical lenns such as (9.102)
(9.108)
Then, the X(1r!2) pulse will produce e(T+) "" - I.:T cos DT + IzT sin DT
3.
- 2'flI2[(IxIIy2 + I y 1Iz 2) cos DT
which are just whm one would have for .a pair of coupled spins. It is straightforward to evaluate the commutation for a specific choice of H 12. Thus, for dipolar coupling, we get
[1i12,(IyJ + [y2)] = BI2 [2Izl I z2 -
1:0:1/2:2 - [yl[y2,(1y1
+ [y2)]
= B I 2[2(Izl,!yl)Iz2 - Uzl,Iy d I :t2
+ r)J 2
(9.103)
From (9.103) we see that this commutalOf at this stage has one-quantum coherence. Indeed. we can see this in general by calculating a matrix element of
[H u ' I yT ] (Ma,!,H s8 I y'l'
"" L:
-
IyTHuIM'a')
(9.104)
0"
(9.105)
so that
I.)It2] sin Dr} ..
(9.110)
8S '
['H s8 , IyTll
(9.111)
Here we will have teons in the double commutation '
(9.112)
[Hjk> ['Hlm,(IYIII + I yl )]]
(9.113)
Now, if neither j nor k is 1 or m, the outer commutator will vanish. Thus, we get either two-spin terms such as
Thus M and M t are states which are joined by IyT' But IT)
- Ii) +ui -
DT
but since either for m must equal n, this will be of the foon
I
"" L:(M al'HsslMa")(M a"IIyTIM 0")
~i([,t -
II Ii) cos
We see clearly the one- and two-quantum coherences. The fact that the B I2 tenn is proportional to T is analogous to the small r behavior of the tcnn sin (aT!2) of lines 3 and 4 of column 10 of Table 9.1. Clearly, we can use a similar approach to the next Icon in the series of (9.100)
[Hjk, [Hln!' [ynn
which, using (9.97), which says 'H S8 is diagonal in AI, gives
Iy'f ""
+ [I.: lui
'12 (.I,: )' [H
(Mal'Hs8IM"a")CM"a"IIyT1M'a')
/If'' ,a"
- (M alIyTIM' al/)(M'a" IH8IM' a')
Expressing the [z and I y operators in tenns of raising and lowering operalOrs, (or using Table 9.2) we have
3i T{ + + ~ 2'Bt22i (II I 2 -
= -3iUz1 l z2 + 1:1 [z2)BI2
3.
(9.109)
e(T+) "" - ItT cos DT + Ix'!' sin DT
+2(Iz2.Iy2)Iz1 - (lzz,!y2)Iz d =-"2BI2[(I\+ +[1- )Iz2+ Jzl (1+ 2
+ (Itl I y2 + I y1 I z 2) sin Dr] ...
(9.106)
['H 12' ['H 12,Uy l + I y 2)ll
(9.114)
or three-spin terms such as
M""M'±l '60
(9.107)
(9.115) 461
In this mannerlwe see that each higher order or T in the series (9.IOO) adds one more possible spin to the couplings. We can now see another userul point by looking in detail at (9.103). AI· though lhe top line involves products of three spin operators, the commulation step reduces the prodUCI to two operators (e.g. ZI with [22)' In a similar way, one finds that the three·spin tenn (9.115) involves three spin operators, such as [zll1/2Iz]. In general each higher tenn in the series involves PrOOtiCts of the previous tenn Wilh a pair of spin operators, bUI lhen the commutator reduces the num~r by one. Thus, each successive lenn has one more spin operator in the product. Since it takes three spin products to get a three-quantum coherence, the T 2 lenn is the first one in the series which can give three-quanlUm coherence. The T] tenn is the first one which could give four-quantum coherence. Note that we say "could", implying that il does nO! necessarily do so. Thus, if one has two spin-j nuclei, the highest-order coherence one can produce is second, hence the T 2 , T , and higher tenns cannOI produce more lhan second-order coherence (double-quantum coherence) in this case. To find what coherences are aClually realized, one mUSI look in detail at the lenns or use the general rules about the maximum coherence 2NI of a group or N nuclei of spin I. Suppose in (9.115) one has a three-spin prodUCI involving coordinates of spins 1,2, and 3. Expressing the 11:1: and lyA: 's in tenns of raising and lowering operators, one might ask, will there be tenns in ()(T-) Uust berore the second. X(1fn) pulse] such as
I
ItriI:
?
(9.116)
The answer must be "no" since this is a three-quantum tenn whereas we have said each tenn in the infinite series prior to the second X(7fn) pulse has one-quanmm coherence. Permissible terms then would be of the general fonn
IiliIt
or
(9.117a)
It 1:2[,]
(9.l17b)
These may be thought of as grouping a raising (or lowering) operalor (e.g. with either zero-quantum operators or with paired raising and lowering operators (e.g. Ii It)· When these are operated on by the Zeeman operators. they simply multiply or I; by a phase exp(inT) or exp(-Wr) respeclively, or leave the the I:k's alone. Thus. the second X(7rn) pulse still acts either on expressions such as (9.117a or b).
Ii)
I:k
r:
Now
xIi X: X(l~t + i/y 1)X: 1",1 X/ 1X: /"'1 + il,l
so
il:1
X(It Ii !jJX = (1z1 - iI: I )(lz2 + il:2)(I",3
+ i/", I 1:2 /",] + 1"'11:21:] - iI: 11",2 1 z3 - 1:1 1z2 1:] + J:t /:21~3 - il:II:21:3
.= 1",1/",2/",3 - i/",t 1",21:3
(9.118)
When we express [zl> 1",2, and /",3 with raising and lowering operators, we see explicitly how the three-quanlUm coherence arises. TIle lenn (9.ll7b) will give
XIi h:IJ:X
: (lzt - iI,t)ly21yJ : [.1;\ I Y2I y] - il: 1l y2 / Y:1
(9.119)
TIle tenn l"'t1Y21Y3 contains three-quantum coherence. From our analysis of the power series expansion, using (9.9a), our con. c1usion is. then. that the pair or pulses X(1:(l) ... T •.• X(7(fl) will generate all orders of quamum coherence permitted. While it is a question of proper choice of T to achieve lhe desired orders, in general, all allowed orders will be presem. We thus have IwO problems yet to consider. (I) How can we detecl a particular desired order? (2) Is it possible to generate a particular order without generating other orders? We tum to these IOpics in the next two sections. 9.3.5 Selecting the Signal of a Particular Order of Coherence We turn now to a discussion of how [0 select the signal arising from a panicular order of coherence. We have already mentioned one approach at lhe end of Sect. 9.3.3, where we noted that if the signal is Fourier analyzed with respect to the evolution time t], the coherence or order p gives rise to lines at frequency pil, where is lhe amoum one is off resonance. Thus. introducing a deliberlarge compared to the spin-spin splillings ate resonance offset by an amount produces groups of lines in which the spectra of the different orders are well separaled. This approach, and a variant or it (lime proportional phase incrementation, TPPI) in which the offset is actually zero but is made 10 appear nonzero, has been widely employed. especially by the Pines group [9.5,7J. We discuss TPPI at the end of this section. AnO!her powerful method was imroduced by WotOIUi and Ernst [9.I7J. We turn to il now. Their method is based on a recognition that the phase of the multiple quantum coherence depends on the phase of the exciting pulse in a manner which depends on the order of the multiple quantum coherence. Their concept also underlies the scheme for production of a selected order of coherence which we take up in the next section. The basic Wokaun-Emst scheme may be described as follows. Let us denote by Xf>' a 7(/2 pulse applied about an x'-axis in the rolating frame where the Xl. axis lies in the x-y plane. making an angle ¢ with the x-direction according to lhe equations
n
x' "" x cos ¢ - y sin ¢ 462
- il:])
n
(9.120.)
(9. 120b)
I;,:I = I;,: cos ¢ - I y sin ¢ = eil.t/J I;,:e-il.t/J
(9.12Oc)
Note Ihal Xl is rotmed from x in a left-handed sense about I z by an amount Wokaun and Ernst apply a sequence Xt/J ... r ... X",,,.tl ... X ... t2
1J.
(9.121)
recording the complex signal 5(T, 1>,tl, t2) as a function of t2, for fixed values of the parameters T and tl, and for a succession of appropriately chosen phases 1>. Let us write the signal for ¢ = 0 as a sum of contributions G p from the various orders of coherence p (p = - N 10 +N)
1 I
Then, utilizing the fact thm X = eiJrTff/2 we have Xt/J = exp [i!(l;,:T cos ¢ - I lIT sin
L
G"
-!...
G =
h
P
J
5W,e-; pod.
.
(9.123)
o
Since in practice Ihere are 2N + I values of p, measurement of 5(1)) for 2N + 1 values of 1> should suffice, as we show below. To understand the Wokaun-EmSI theorem, we need to
1. 2.
examine Ihe general form of the contributions of a given order of coherence, examine the effect of a phase shifl on the differenl orders of coherence,
,nd 3.
show how to ulilize Ihis knowledge to pick OUI the contributions G p of the particular order of coherence. We consider a system characterized by a Hamiltonian in the rotating frame 1{ =
-
E hDkI:k + E k
1{jk
==
'Hz + 'H u
(9.124a)
X", = exp(iRI;,:TR- I 1r!2) = ReilrT1r/2 R- 1
= eil.Tt/J Xe-il.Tt/J
where for the spin-spin interactions 'Hjk we keep only the secular terms so Ihal (9. 124b)
Then, we have the quantum numbers NI, a such that (9.1240)
'HuIMa) = EnlMa) Note Ihat (9.124a) allows for a variety of chemical shifls. 464
(9.125d)
We now wish to express the ¢ dependence of the signal, 5. Abbreviating 5(T, 1>,tt> i2) as 5(t/J), we have S(t/J) = Tr {
If [e -(i/h)'H"J2 X e -(if h)'H"tl J£:I(T+, t/J)
[inverse]}
(9.126)
where (I(T+,t/J), defined as f!{,,+, t/J)
== Xt/Je-(i!h)'HT X",IzTX;le(i/h)'HT X;I
(9.127)
is the density matrix jusl after the second 1r/2 pulse, for the case that both 11' pulses have phase shifts t/J. The tP dependence of S(t/J) therefore arises in £:I(T+, ¢). Expressing Xt/J utilizing (9.125d) and utilizing the faci Ihal I zT ' and thus exp (iIzT¢), commutes with H, we get f!{T+, t/J) = eil,Tt/J Xe-(i/h)'HT X IzX-leCi/h)'HT X-le-i/,Tt/J
= eil.Tt/J £:I(T +, O)e-il.Tt/!
(9.128)
We recalllhm (1(,,+,0) is the density matrix just after the second X pulse, hence it contains all the multiple quantum coherence. In general, for a given spin system, £:1(,,+,0) consists of a number of terms representing not only the different possible orders of coherence, bUI also the various ways of generating a one can generate p = 0, given order. Thus, for a three-spin system, of spins ± 1, ± 2, ± 3. Consider p = I. It could arise from a variety of operators such as
!'
k>j
['Hjk> IzT] =0
(9.125c)
we gel
p=-N
,.
(9. 125b)
R=ei/.Tt/J
(9.122)
and then consider 5(1J) to be known as a continuous function of 1J. The essence of the Wokaun-Ernst method is then that the Gp's can be found as transfonns of 5(.)
1»]
Defining
+N 5(.=0)=
(9.125')
It
It12:
It h.hz
ItIiIi
(9.129)
plus all the operators one can get by permuting the labels 1,2, and 3. Utilizing a symbol f3 to designate all such various ways of generating a given order, we can clearly break £:I(r+ ,0) into a set of components Yp{J of given p
"T+,O) = Lgpp p,p
(9.130)
465
Now, a given order, p, is distinguished by having p more raising operators than lowering operators. (If p is negative, there are more lowering operators than raising operators.) We can then easily evaluate g(r+,4J) as i ' ,1'.;P9p13e -i/,T.;P (9.131) e< r +, f) "" eiT'TtI> g( r+, O)e -it ,T.;P =
l.:e
P.P
But,
ei/'T h±eif,.;p ""
h
±e±i.;p
eif'T Izke-i/'T "" Izk
by p, and in which the coefficients of the hannonics are the Gp's. Therefore, if one knew S(tjJ), one could deduce the Gp's by taking the Fourier transfonn of 5(<1). Thus, suppose we have deduced S(f). (We can do this by measuring at a number of discrete values fi and using a smooth interpolation procedure to estimate S for values between adjacent pairs of values fi') Then consider I, the transfonn of S, 2.
(9.132) 1=
Taking expressions such as (9.129), inserting factors exp (-ilzTf) exp (ilzTf) between each pair of spin operators, and using (9.132), one gets results such as
e i/'T4>
J5(
Substituting (9.1 36a) for S(f) we get
It h ze- iT""4> "" It h ze i4>
2.
I = L:G p
(9.133) That is, the effect of the f rotation for any operator with p "" 1 is simply to multiply that operator by exp (if). In a similar way, the rotation operators acting on a p "" 2 operator merely multiply it by exp (2if). In general, for the operators
= 21rG
I
p
+ '" G p [e 21l"i(p-p') - I] ~ itp' p) p f. pi
= 21rGpl
9p/3'
eif,T4>Yp/3e-i/.T4> "" 9p/3e iP4>
t
11<,+.<1)=
p=-N
(9.134)
fl· (9.135)
(L:9pp)e;po /3
where N is the maximum allowed value of p for the particular spin system. Substituting this expression into (9.126) we get the signal as a function of f to be 5(<1) = pEN e;poT{r:t [e-O/")"" Xe-Olh)'Ht, ] (1gpp) [iowse]} N
"" L
G p e iP 4>
where
(9. 136a)
p=-N
G
p"" Tr{ r.t
[e-(i/h)"Ht 2xe-(i/h)1"f.t l ]
(~yp.B)
[inverse]}
. (9. 136b)
The individual tenns 9p/3 depend on r, on the ilk'S and the strength of the spinspin couplings as illustrated in the tables we have worked out for g(r+) for the two- and three-spin cases. Examination of (9.136a) shows that it says that S(tjJ) can be represented by a Fourier series in tjJ in which the harmonics are labeled
466
Je;(p-p')Od 0
p
Note that this result is independent of We can utilize (9.134) to get
(9.137)
0'
(9.138)
2.
Gp =.2.. 2.
J5(
(9.139)
o
In some systems, such as a dilute solUlion of some molecule of interest, there is an upper limit to the quantum order [the quantity N of (9.135)]. In Olhers (for example an effectively infinite solid of dipolar coupled nuclei, such as CaF2) N has no limit. Suppose we had the fonner case. Then, there are 2N + 1 values of p, consequently 2N + I Gp's. Clearly 2N + 1 measurements of S(f) should suffice to deduce the Gp's. Suppose, then, we pick the 2N + 1 values of f, which we label fk' as
2.
= '2N + 1
(9.140.)
k=0,1.2..... 2N
(9.140b)
Note if we chose k "" 2N + I, we would have a value of f "" 21r, which produces the same shift as ¢ "" 0, a value we have already obtained from k "" O. Then 5(<1,) ,,5, = L:Gpexp(ip
+N =
L:
Gpexp [ipk2p/(2N + 1)]
(9.141)
p",,-N
This is a finite series. Consider then the transforms
467
I:S(1,)e>P (-ip' 1,) k
L: G e>p [i(p - p')1d k,p " = L: G p L: "p [i(p - p')kh/(2N + I)J
5(0)+5('11") = L:Gp(eipO+eip1r)
=
p
p
(9.142)
p
=
Now, defining
2N
=
1.=0
1.=0
1,,=0
=
=
(
L:I'
)
L:
I' I
f2N+l
-1-1
G 1(1 +ei1r/2 +ebr +e i31r / 2)
if
1. 1
(9.144)
1,,=0
Examination of (9.144) shows that as long as p 1= pi the exponent in neither zero nor a multiple of 21l', hence
f
f
is
=F 1. But
(9.145)
f2N+l = e i (p-pl)211" = 1
I
p
=
L. '1'1. e -;P'¢' ( - -I - )"'5(")
2N+l
with
,
(9.146)
k=0,1, .. .,2N
Therefore, measurement of S(¢k) for these 2N + t values of
to decompose S into the contributions. Gp , from the various orders of coherence. Since Gp is a complex number, one must record signals S(r/J) as complex numbers, Le. in quadrature, to evaluate the fannula of (9.146). It is sometimes useful to carry out a simpler version of adding signals. One picks a set of phases (e.g. O. rr(2. '11". 3'11"/2) and then a~ds the signals together. Sometimes one adds signals of one phase but subtracts signals of another. To see how this works it is useful to keep in mind a graphical picture. From (9.136a) we have 5(1) =
+N
L:
Gpo;P.
(9.147)
p=-N
hence each order can be represented in the complex plane by vectors exp (ip¢). Consider then an example with ¢ = 0 or '11". Then 468
(9.150)
The four complex numbers may be thought of as four vectors in the complex plane (1. i, -1. -i) which. when added tail of one to head of the prior one, form the sides of a square, giving a zero result. Thus there is no signal from Gt, first-order coherence. Likewise. there is no signal from p = 2 or p = 3. But for p=4weget
(9.151)
k
1k=k2,/(2N+I)
(9.149)
G4(eiO +ei4..-/2 +eiS..-/2 +e il 2..-/2)
hence the sum over k in (9.144) vanishes. If p = pi, f = 1, and the sum over k is trivially 2N + 1. Thus
G
L Gp(eipO + eip1r / 2 + eip1r + e ip3 '1f/2)
Clearly for p = 0 this will give 4Go. For p '" I this will give
k=2N+ 1
(1_1"'+1)=
(9.148)
p
00
= L: I' = L: I' -
Gp
even
5(0) + 5(,(2) + 5(,) + 5(3,(2)
p')k2,/(2N + I)] 00
Gp(l - I)
p odd
Thus, all the odd-order coherences contribute nothing. If we lOok ¢ = 0, rr(2, rr, 311:/2 and added signals we would get
we have
2N
L:
2
L:
GpO + I) +
even
p
(9.143)
f.:: exp [i(p - p')21r/(2N + 1)]
L: e>p [i(p -
L:
=
,
That is, all four terms exp (i4¢) are equal 10 I. Thus we will record Go and G4, but not Gj, G2. or G3. In a similar way one can show that we will not record Gs, GG, G7. but will record Ga. In fact. as we see from (9.151), we will get a nonzero result whenever all the tenus are of a given order are 1. Since for a phase shift increment of 8
L Gp[l + eip8 + (e ipO )2 + ... J
(9.152)
p
we will get signals when e ip8 = 1 or
pIJ=kh
,
k=0,1,2,3
(9.153a) (9.153b)
Thus, if we get a signal for p = 2tr/O, we will also get a signal for 2p, 3p, etc. One sometimes speaks of a sequence as being "pk selective", since if it gives order p it will also give k times any integral value of k. Note that as long as we add all the signals, we get Go. the zero-quantum signal. To eliminate it we add half the signals and subtract the other half. Some cases given by Wokaun and Ernst are
469
add add add add add
q, =0, 1f gives p= O. 2. 4. 6, 8 q, = O. subtract; = 1f, gives p = 1,3.5,7,9 q, = O. rn. 11", 311"n gives p = 0, 4, 8, ... q, = 0, 1f, subtract; = -rr/2. 3rrn.., gives p = 2,6, q, = 0, 2-rrf3, 4rrl3, subtract if> = -rr13 + (0, 2rrf3,4rr(3), gives p = 3, 9.
These simple sequences are particularly useful under circumstances where the allowed values of p are limited. Thus, if p = 9 can be ruled out, the last sequence gives one just p = 3. We remarked mat one method of displaying spectra by quantum order is to tune off resonance by an amounl Then lines associated with order p can be identified by Fourier transforming the signal with respect 10 Ihe evolution time tl, since they occur at a frequency offset of pil. As we remarked, there are disadvantages with working off resonance. A solution is to use a technique called time proportional phase incrementation (TPPI). TIle idea is to work at = 0, exact resonance, but in collecting dala for the variable t 1 (the evolution lime) to introduce a phase shift which depends on tl. If the phase and tl obey a linear relationship
n.
n
(9.154) Then as tl is advanced the extra phase is just as though one were off resonance by an amount Llw given by Llw= A
(9.155)
Then, when one does the Fourier transform of the signal with respect to tl, an effective frequency offset obeying (9.155) is introduced. This approach is unaffected by inseniOIl of 'II" pulses to refocus magnetic field inhomogeneities or real frequency offsets aI any part of Ihe cycle (Le. al me midpoint of the preparation cycle, the evolution cycle, or the detection cycle).
9.4 High Orders of Coherence We have seen some simple ways of producing and delecting multiple quantum coherence. For many purposes lhese memods suffice. For example, if one wished to distinguish protons in a CH3 group from protons in a CH 2 group. one need only to be able to produce and detect two- or Ihree-quantum coherence. What happens if one is concerned with much higher orders? II is easy to see that there is a fundamental problem. Each quantum order contains "information" - that is. it represents order in the statistical mechanical sense. But we are confident mal the tOlal slatistical mechanical order will never exceed its initial value representing the sample's initial magnetization. If one pumps only loworder quantum coherences, the statistical order is shared by only a modest number of (MaleIM'aI)'s. However, if one sets the time, T, of the preparation to longer
and longer times one finds one exciles most of the low-p elements of e, as well as matrix elements of progressively larger p. Thus, the stalislical order is being shared by more and more elements of e. As a result, individual elements gel much small~r. One lends to lose in two ways. (I) During preparation one spreads the order among many different elements of f!, and (2) the detection pulse feeds back many elements (Malf!IM'o') in which though M - M' = I) may be the same, there is desu1Jctive interference from the range of values of the other quantum numbers, a. It is thus clear that 10 get strong signals one should first of all restrict the elements of f! which one excites to just the desired ones. For example. if one is interesled in the six-quantum spectrum, one would like to excile only the sixquantum coherences. Then. having excited the desired coherences, one would like to read them out at a later time in a manner which avoids any further loss of coherence. We turn now to these mailers. Many of the ideas involved here are from the studies of Pines and his students. The two review articles by Weilekamp [9.5] and by Mwwwilz and Pines [9.7b] are especially useful for study, as are the basic papers by Warren el a!. [9.18.19) and Yen and Pines [9.20).
9.4.1 Genef'::.ting a Desired Order of Coherence Can one generate one and only one desired order of coherence? We have seen that it is possible to delcct an order 1J and integral multiples of p by cycling the phase of the exciting pulses relalive to the detection pulses. We will now show how to use phase cycling to generate coherence of a given p and multiples thereof. FIrsl. however, it is useful to introduce anOlher way of looking at generation of multiple quantum coherence. Utilizing two pulses. we have seen how the firsl 1fn. pulse convens the density malrix from its initial value, I:, to I z , producing one-quantum coherence. As time develops, spin-spin correlations develop in lhe density matrix as is evident either from the spin operalor fonnalism [e.g., (9.90)] or from lhe perturbation expansion (9.100). However, the density matrix still possesses only one..quantum coherence. The second 1fn pulse, by acting on onequantum terms like I:II:2It which express spin-spin correlations (in this case of three spins) produces multiple quantum coherence through operators such as We can express mis fonnally by writing the equation for e(T+), where T+ is the time just after the second 11"/2 pulse. For convenience we will take the second pulse to be an X, though the argument we present could be done, a bit less elegantly, with an X pulse. Since X undoes an X pulse. we have
It Ii It·
X = X-I e(T+) =
so
X-I
ex p (
(9.156)
-*1iT)X I:X- (*1iT)X I
eX P
(9.157)
Suppose for concreteness we are exactly at resonance, with 1{ being only the spin-spin coupling. We take
470 471
(9.158)
1t = 1t zt where 1t zz could be the usual secular dipolar coupting, such as
1t zz =
I::
Bjk(l~jI~k + IyjIyk - 2IzjIzk )
(9.159)
j> k
Then X-I ex p (
-~1tzzr)x
= exp [
_~(X-l1tHX)T]
= ex p (
~~1tyyr)
(9.160)
coherence?" In fact, as we shall see. we can find a way to generate coherence of order pk. where k = 0,1 •.... By special means, one can eliminate k = O. Then. the means to distinguish among the other k values is by the duration of the e?tcitation period. How then do we limit generation to an order pk? We will find that there are two parts to the task. The first part is concerned with finding a way to limit the order, and the second part. with making the amount generated large. In order to keep the discussion sufficiently general, we will therefore not as yet specify the Hamiltonian. However. we suppose that as with 1t yy it can generate various orders of coherence, p. That is
so that
'Ii =
e(r+) = exp (
-~1tyyr ) I z exp (~1tYYT)
L
(9.166)
p
(9.161)
where. for example. if 1t zz were given by (9.159),
1tyy =
I: 'lip
BjkU~jI~k + IzjIzk - 2Iyj I yd
(9.162)
j>k
We now want to see how to limit the orders generated. We already have a clue from the ideas of Wokml1l and Ernst; we should utilize the phase shift properties. Let us think about what happens to H if we shift the phase of the rf pulses which generate it. For the case of two rr/2 pulses so that the coherence is generated by
In this formulation, explicit mention of the pulses has disappeared, replaced by a transformed Hamiltonian. 1t yy , which starts to act at r = O. Now, e(r+) possesses. as we have seen, multiple quantum matrix elements. That is
xex p ( -*1t.ur)x = ex p ( -*1tyy r)
(9.167)
we have seen thai (9.163) X<,6ex p ( -*Hztr)Xt/>
However, since the density matrix at t = O. I z , has no matrix elements between different stales M and M', we may say that 1t yy acting over time r generates the multiple quantum coherence. Indeed, expanding the exponentials we have exp ( -*1t yy r ) = 1 _ i
1t~yr + (~y 1tyy ~yyr2
+...
=
(9.164)
x
From (9.162), substituting raising and lowering operalors (9.159).
1tyy =
I::
(9.165)
which therefore contains zero- and two-quantum coherences. We have secn numerous examples of ways in which the ingenious resonator can effectively modify a Hamiltonian: spin echoes. in which magnetic field inhomogeneities are effectively reversed; magic echoes, in which dipolar dephasing is undone by creating the negative of the dipolar Hamiltonian; and spin-flip line narrowing, in which the dipolar Hamiltonian ?i zz is averaged to zero by being made to jump to 1t~~ and H yy . Therefore, we may rephrase our question "can we generate one and only one desired order of coherence" to "can we find a way to achieve a Hamiltonian which will generate one and only one desired order of 472
eXP(*IZT~)X ex p ( -*IZT~)
=
eXP(*IzT~)X ex p ( -*1t zz r)x ex p ( -*IZT~)
=
eXP(*IZT~) exp ( -*Hyyr) ex p ( -*IZT~)
Bjk(t(ItIt" + IjIi;) + IzjI zk
j>k - tut Ii; + Ij It»)
eXP(*IZT~ )X ex p ( -*IZT~) ex p ( -*1tztr)
(9.168)
Let us then think about a general Hamiltonian, which for zero phase. ~, we call
1to. We decompose it into various orders of coherence (9.169a)
Then 1tt/> == e i1•T <,6Hoe -il.Tt/> =
I:: eil.T¢J1tpe -il.Tt/>
(9.169b)
p
In the spirit then of (9.133), we then get
473
L 'Hpeip
'H
(9.176.)
(9.170)
p
the complex numbers of (9.175) fonn a closed polygon of exterior angle P
This expression is reminiscent of (9.136) S(,,) = L: Gpe;P.
eip,p = 1
(9.176b)
p
Therefore, if we pick a p, which we call ]Jo, that we wish to generate, we just need to select
The way we got selectivity in signal detection was to form a sum of signals for a sequence of phase shifts. How can we use this idea for Hamillonians? We have already found out how when we studied spin-flip line narrowing. What we must do is employ the conditions for a time average Hamiltonian. That is, we switch among the various values of the Hamiltonian, fonning a cycle of desired values, then repeat the cycle again and again for some desired total time. The development of the system for one cycle in which the Hamiltonian takes on n values 1=0 to
'HI,
(9.171)
n-l
each of duration Tk> is given by the operator exp ( -*'Hn';i
exp
1'1",,_1 ) .•• exp (
-*
'H2T2) ex p ( -*'HOTO)
(-2.t, L'HITI)
(9.172)
n = ]JO
(9.177b)
we will have a non vanishing 'Hay' We will also have a nonvanishing 'H. for p's which are integral multiples of po. Thus, we will select values p
l
p=pok
k=0,1,2
¢ = 2'11"/vo
using
(9.178)
For all other values of p, (9.175) and thus 'Hay vanish. The theorem we have proven holds true for any Hamillonian provided we satisfy the condition for average Hamiltonian theory:
(9.179)
1
10 = L:TI
(9.177a)
which from (9.175) gives
1
where the equality holds provided the Tk are sufficiently short. Then the average Hamiltonian 1 'Hav = t L 'HITt where (9.173a) o
VOr/> = 2'11" or
(9.173b)
The sons of correction tenns can be looked at using the Magnus expansion (see Appendix K, and references therein). Reference [9.20] discusses correction tenns in great detail. If we now look at the sequence of a pair of '11"/2 pulses which we have been considering, we have for 'Ho
'Ho = 'H yy
(9.180)
1
So, if we picked a set of n phases
(9.174.)
=
~ L: H p
p
("f:1e;P11»
(9.174b)
1=0
where in the last step we took
+ eip
We have seen in the previous section [note in particular (9.150)] that if 474
(9.175)
From (9.165) we see it contains only the tenns p=-O and ±2. Now, we know that the two-pulse sequence can generate all values of p, but those larger than two require the T 2 or higher tenns in the expansion (9.164) of the exponential. These are the tenns we do not wish to have since we require in (9.174) that the average Hamiltonian be valid. The question is, then, what can we do to create an 'Ho which gives high orders of p without violating the conditions for average Hamiltonian theory to hold? The solution was found by Warrell et at (9.19,20]. We tum now to an explanation of their approach. The Hamiltonian 'H. yy resulted from the action of the pulses X and X(=X-l) acting on 'H u . The crucial steps are (9.160). In panicular,
'H yy =- X-1'HuX
(9.181)
Of course, we start with 'H u - that is what nature gives us. We get 'H yy by 475
applying a transformation (physically realizable!) with the operators X-I and X. Let us then look for some other operator - call it R - to use instead of X. That is. ?-lo is to be produced by applying R to HH' Thus ?-lo:: R-1HHR
so that
exp [ - (Vli)'ltot] = WI [exp [ - (i/li)'It"tJ] R
(9.182) (9.183)
What properties should R possess? The trouble with X is that i( produced solely z.ero- and two-quantum coherences. What we need is an 1-lo that is rich in the values of 1) which we desire to excite. That is, in (9.169a) we want large Hp's for (he values of p we desire. It is clear that no simple rotation operator like X or Y will do, since it simply rotates the basic dipolar coupling. However, we remember that actually
of producing -?-lzz (actually -'Hzz(l). The first method was used to produce magic echoes in Sect. 8.8. It involved producing a strong HI in the x-direction in the rotating frame which one phase shifted by 11" at periodic intervals. Several other'methods involved trains of X and X pulses, as for example shown in Table 8.4. Another useful sequence produces 1{n = k(1{yy -
R:: exp [- (i/h)Hntp]
(9.185)
where 1-l n is the Hamiltonian which acts during the time t r . Now, ?-In can be a real Hamiltonian, such as is the case when R:: X, or it can be some effective Hamiltonian created by the experimenter by manipulating a real Hamiltonian with pulses. The game, then, is to try to guess an H n which does what one wants, then, if it does, figure out how to create it. Note that in order to create both R and its inverse R- 1 , we must be able to create both Hl/ and -Hn. There are not many choices. We have already excluded Hn's which are proportional to I z1" IYT' or I~1" All that is left is Hn :: 'H zz
(9.186.)
'Hn:: 'H yy
(9.186b)
?-In = 'HH ?-In:: ?-l~I~1
01
(9.186c) (9.186<1)
where z' is an axis other than x. y, or z. We could also take linear combinations of the operators. or take these operators multiplied by a constant ?-l~~ alone is not useful since in (9.183) R would commute with the exponential, leaving ?-lo :: ?-lH' However. ?-lu, ?-lyy, or ?-l~I~1 all will do the job. Before showing this, we recall that we get ?-lu; or ?-lyy simply by sandwiching ?-l~~ between two 1r(l pulses (Y and Y for ?-lu, X and X for H yy ). We have seen several ways
•
(9.187)
which, for dipolar coupling given by (9.159), is Hll ::
L
Bjk(IzjIzk - IyjIyk)
j>k
1
::"2 L
(9.184)
where H I is the strength of the rf field, and II' the pulse duration. This term is nothing but an approximation to the time development operator for the system while HI is on, an approximation bec
7-{zz)
Bjk(It
rt + Ii Ii:)
(9.188)
j>k
This sequence arises from the sequence of (8.207). which we rewrite as
(M. X, d,x, Ll,X. d, X, Ll,X. Ll',x, Ll.X. d, X, iJj2)
(9.189.)
where
d::2Ll+t p
(9. 189b)
and t p is the pulse duration [9.19). The fact that Ll' oF 2Ll is to correct for finite pulse length. Note that since (9.188) is a two-quantum operator. we can easily create its negative by phase shifting the pulses by 11"(2, i.e. by replacing X and X by Y and Y. In practice the schemes utilized are all based on sequences of 1r(l pulses. We now want to show that if we have such a dipolar Hn we can generate an Ho which is rich in large quantum orders HJI' Referring to (9.182) and (9.183), we have
11.0 :: R-1HzzR :: exp [+ (i/h)HnT]Hzz exp [ - (i/h)7tnT]
(9.190)
where T is the length of time we let the system evolve under Hn or under -1{Il· Note that in the case 'H1l:: 'H zz , we produce a negative (-Hn) of the form -H u {2. which requires that we let it act for a time T' which is twice T. However, if we set T':: 2T, (9.190) still holds. We now expand the products using the relationship [9.16} e"Be-"=B+[A,Bj+[A,[A,B])+...
(9.191)
Substituting this into (9.190) gives
?-lo
=
iT 1 T2 I H zz + "h['H n , 1i~~] + '2 2T h2 [HII, [Hn.Hzzll + .
(9.192)
Evaluation of the commutators is straightforward but tedious. The general charac-
476 477
ter of the result is clear. however. Successive terms involve progressively larger numbers of spin products, hence include terms which correspond to progressively higher orders of quantum coherence. That is, they contain t~rms of progressively larger p. If T is very shon. only the first few terms. con~bute. and ?~ly small values of p result. But if 11lniT ~ I. many terms WIll be Included, gIVIng large p's though each individual term is still of order 11l.uJ • We now have the operator If we define a time L\T' for which zz alone acts, and a time L\T for the time in the cycle for which we have a fixed phase.
llo.
ll
¢,. (9.193)
LlT=T+T+LlT'
-*ll
U
L\T
1
)R
(9.194)
~=h
(9.195) Po For one cycle. ¢t will then, according to (9.175) and (9.195). need to take on the different values (9.196) ~l =0, (2./Po), 2(2'!l>o), 3(2P//>o) ... 0>0 -1)(2·/Po) This cycle lasts for a time t c given by the number of phases n limes the duration of each phase
=nLl, =PoL\T
(9.197)
However, in order for the average Hamiltonian theory to hold. we must be able to approximate exp (
-i1itfl,o_tLl,')
eXP[-*(1iI/>,O_1 +
ex p ( -i1ioLh') 1
+ 1iO)LlT
]
by (9.198)
This condition is satisfied if
l1ioiLlT' <: I
(9.199)
Thus. it is not the long time L\T which comes into satisfying the Magnus expansion. but the (shorter) time LlT'. Since 1io is independent of L\r'. we can set T and T' (if T' =f:. T) to make 'Ho what we desire. then set Ll,' to satisfy the Magnus expansion. 478
3. 4.
The total time to complete (1)-(4) is PoLlT, and constitutes one cycle. The average Hamiltonian produced by this cycle is I
Ll L 'H p
(9.200.)
p
T
where the prime on the summation indicates including only those p's for which
Suppose then that we wish to select I' = Po. From (9.178) we see we must picle a phase shift 4J for (9.175) given by
tc
2.
.We apply 1in for time T. 1iu for time L\r'. then -Hn for time T (or 7"). This takes time LlT. With the phase of 1in shifted by 2Jr/Po. we repeat step l. With the phase 1i n shifted by 2(21fIPo) relative to step I, repeat step 1. Continue the phase advances until ¢ = (Po -l)(2'11"IPo) relative to step one.
'H" =
= ex p ( -in-l1iunLlT')
=exP(-i1ioLlT')
I.
LlT'
But the development of f! over the total time LlT is given by R- 1 exp (
We may describe the total preparation period to generate coherence Po (and integral multiples thereof) as follows:
k=0.1,2, ...
p=kvo
(9.2001»
Note that we multiply the summation by L\T'/L\T since during each interval LlT. 1i1/> is acting only for the time LlT'. The rest of the time (T+ 7") is spent generating Rand t. The basic cycle is then repeated for as many times, m, as is necessary for the amplitude of the desired coherence to grow sufficiently. Thus the total preparation time 'p is mLlT. To give an idea of what is involved, we may quote the numbers reponed by Warrell el al. [9.18] to produce a selected fourth-order coherence for protons of benzene molecules oriented in a liquid crystal. They use the sequence (9. I 89a) to produce
n-
1io=
n- l 1iu n
(9.201)
with a 2.4 lis duration of X (or X)
,=3.0ps
,
2T'=8.8p.s
•
T=2ms
,
L\T'=35jts
(9.202)
The entire cycle was repeated four times. To get rid of zero-order terms, one must reverse the sign of for half the values of ¢ in the cycle. For a Hamiltonian such as lhat of (9.201) this can be achieved by choice of the phase ¢, since adding 'Ir/l to ¢ changes lhe sign of 'HI>. We now have a situation in which many more than two pulses are used to generate the lJo quantum coherence. We may then introduce the unitary operator U(T p ) given by
HI/>
U(rp ) = "p [(Vh)'H"Tp )
(9.203)
to describe the development of f! during the preparation interval of duration 'p : (9.204) 479
laJ
'. Ibl
t,
u
"
1t
v
u
t,
'.
t,
v
x
where u is taken to be the Hamiltonian in the absence of pulses. It turns OUt to be conveniem to Ihink of V as ending in an X pulse (Fig.9.6b). We then redefine V 10 make Ihal explicit, replacing V(Tm ) with XV(T m ) so Ihal (Iy ) :::
'. L.Ib
Tr{ IyXV(Tm)exp ( -*1t x ex p (
*1t
u
t l ) U(Tp)I,U-I(Tp)
1 u t l ) V-I(Tm)X- }
-*1i
Fig. 9.6a,b. More general preparation lind dc~eelion pulses represented by unilary operato", U lind V, preparation and mixing times Tp llnd T m , evolution and detection times t 1 and t2 respectively. In (b) we introduce explicitly Illi X(Tr/2) pulllC just before the detection period so that V is defined II.S all the mixing pulses which precede the X{'If/2) pulse
'" Tr{ (X-
We then let the system evolve freely for a time tl. To observe the effects of what happened during fl. we need to mix down to one-quantum coherence. Previously we did this with a third 7r/2 pulse. As we shall see, we can do better by applying a fancy sel of pulses, closely related to those which prepared the multiquantum coherence. Let us call the lolal mixing time T m . Then we represent mixing by a unitary operator V(Tm ), such thai
::: Tr{ V-I I , vex p ( -*1tu t l )UIzU- 1 exp
e{Tp
+ tl + Tm ) = V(Tm)e(T p + tl)V- 1(Tm )
(9.205)
The preparation, evolution, mixing, and detection periods may then be thought of as in Fig. 9.6. We now necd to investigatc the optimum way to carry out the mixing so that we do not lose Ihe advamages of confining the coherence we have generated to a small number of orders, p.
x ex p
1
IyX)V(Trn)exp (
(*1tu
1
t 1) V- (Tm)}
(*1t
u t l ) }.
(9.207)
To understand the coment of this equation, we evaluate the trace using the eigenstates of the system. Recall that we have M, the eigenvalue of I: T , as one quantum number, and a family of others related 10 the dipolar energy abbreviated by ~. We thus denOle states by IMa). It is sometimes convenient to use a single symbol r or .! to stand for the pair. We define W r then by
E,. = EM~ '" hw r
and then
(l.) •
(9.208a) (9.208b)
L (rJV- 1I, VI.)e-;··I, (.IVI,V-'lr)e-C"I,) r,'
9.4.2 Mixing to Dctcct High Ordcrs of Coherence
r,'
Our task now is to find an appropriate set of one or more mixing pulses to provide large detected signals when we have generated large-order coherences. The Pines group has done extensive work on this problem [9.20]. Specifically, given a preparation specified by U(Tp), whal should we conSlJUCt for V(Tm)? We of course evemually detect the signal as a funclion of t2, but in order 10 study the nature of V we shall focus on t2 '" O. Thus. if we were going 10 delect (Iy) at t2 =0
u t l ) U(Tp)I:U-I(r p )
(9.209)
Let us first look at a simple limiting case: what we get if both U and V leave (! alone. This would be the case if there were no spin-spin coupling (isolated spins) and no rf pulses Other than the final X pulse. Then U :::: 1, V ::: I. Since the states Ir), Is) are eigenstates of I the matrix elements are diagonal, with " the result W r • vanishes, giving
(I.).
L I(Mall,IMa)['
(9.210,)
MOt
(Iy ) ::: Tr {Iye{T p + tl + Tm ))
: : Tr{ I
y V(Tm) ex p (
-*1t
u t 1)
x eXP(*1tutl)V-1(Tm)}
480
(9.21Ob)
1 U(Tp)I,U- (Tp)
(9.206)
which is just the well·known result of applying an X pulse 10 an initial density matrix I:T . This signal is, of course, the maximum we can get since it corresponds to rotating the total magnetiuttion perpendicular to the static field. There is no loss in order in that process, hence this is the maximum signal we could get. It is
4.,
conceivable that during the development period, tl> there is some loss of order. Let us therefore next examine the general result (9.209) for the case that we eliminate such losses by making t I = O. Then, going back to the last line of (9.207), we get (lJ/')'I=O as (1:\1)(1 =0 = Tr (Iz{VU IzU- 1V-I)}
(9.211)
Comparing this result with (9.21Ob), which gives us the maximum possible (Iy ), we see that if we make
VU = I
(9.212.)
Le.
v=U- 1
(9.212b)
we will guarantee the maximum signal. Physically this condition states that what U does. V undoes. This statement is clearly a general kind of condition analogous to fonning the conventional echo or the magic echo. We can compare this resull with what we would get if we used U to create the coherence. but only a single 7I:n. pulse for the mixing. Since as we see in (9.207) we have already explicitly included a final pulse X, this situation corresponds to having V = I:
(I,)"
=0
=T'(I,(UI,U-')j
(9.213)
says that both V-I and U "rotate" I z • Unless V-I and U produce the same "rotation" the signal is diminished from its maximum value. In general. the bigger the relative "rotation", the smaller the net result. In particular, suppose U generates panicularly large p's, but Y- t does not. Then U will put the order of the system into large p, but Y will not be able to bring it back down to zero order to pennit the X pulse to conven it to transverse magnetization. The Pines group has made extensive use of sequences satisfying (9.212) V=U- I to generate and then observe large-order coherences. It is the author's opinion that the Warren. Weitekamp, Pines melh<Xi of selective excitation of muhiple quantum coherence can be thought of as a kind of graduation exercise in spin Hamiltonian manipulation. It requires that one be completely at home with the idea that pulses manipulate the Hamiltonian, with average Hamiltonian theory, possess knowledge of the specific kinds of Hamiltonian one can generate and how to generate them. It illustrates then how once one has these details in mind, one can advance to a higher level using the ideas as building blocks, much as in electronics one thinks of oscillators. amplifiers, mixers, pulses and so on, without then being overwhelmed by the details of each circuit.
How does this compare with the result when Y = U-I? U is a unitary transformation. This expression reminds us of the dot product between a vector (1z) and the same vector rotated (by U). It should be less than Tr {I;} or for that matter than Tr{(Ulz U- 1)2}. Indeed
T,{(UI,U-')'j =T,{UI,U-'UI,U-'j =T,{UI;U-'j =T'{I;U-'Uj =T'{I;j
(9.214)
Consider then Tr {lIz - (U I z U-I)]2}. Since Ihis is the trace of the square of an Hennitian operator, it must be positive. Thus
oST, {(l, -
(U I,U-'))'} = T, {I; j + Tt «U I,U-')'} - T, {I,UI,U-'} - T, {UI,U-' I,}
(9.215) Pennuting the order of the operators of the last trace, and utilizing (9.214), we thus get that (9.216) In general, as in the Schwarz inequality, we expect the inequality to hold unless U is the identity operator. Thus. the signal of (9.213) will be less than the maximum. We can think of the signal for tl = 0 along these lines: the expression
(9.217) 482
483
10. Electric Quadrupole Effects
10.1 Introduction So far we have considered only the magnetic interactions of the nucleus with its surroundings. To be sure, by implicalion we have considered Ihe effect of Ihe nuclear charge, since it detennines the electron orbits and where the nucleus sits in a molecule. However, we have not considered any electrical effects on Ihe energy required to reorient the nucleus. That such effects do exist can be seen by considering a nonspherical nucleus. Suppose it is somewhat elongated and is acted on by the charges shown in Fig. to. I. We see that Fig.1O.1b will correspond to a lower energy, since it has put Ihe tips of the positive nuclear charge closer to the negative external charges. There is, therefore, an electrostatic energy that varies with the nuclear orientation. Of course,1 turning the nucleus end for end does not affect the electrostatic energy. Consequently, for spin ~ nuclei the electrostatic energy does not split the ml degeneracy.
-q
+q
(a)
-q
+q
-q
+q
(b)
+q -q
Fig.IO.I. (a) A cigar-shaped nucleus in the field of four charges, +q on the z-axis; -q on the y-axis. The configuration of (b) is energetically more favorable because it puts the positive charge of the ends of the cigar closer to the negative charges -q
1 See references to "Quadrupole Effects" in the Bibliography and the articles by Cohen and Rei! and by Das and lfalin under "Books, Monographs, or Review Articles" in the Bibliography.
•
485
10.2 Quadrupole Hamiltonian - Part 1
(We note that sometimes Poisson's equation applies instead. Some care must then be exercised because we are, of course, interested only in the orientation depen-
To develop a more quantitative theory, we begin with a description in tenns of the classical charge density of the nucleus, (!. We shall obtain a quantum mechanical answer by replacing the classical (! by its quantum mechanical operator. Classically, the interaction energy E of a charge distribution of density (! with a potential V due to external sources is (10.1)
E = j g(r)V(r)dr We expand V(r) in a Taylor's series about the origin:
w)
V(r)=V(O)+LX aO 0' Xa
\
r=O
a'v +21 LXaX/3a a . 0',/3 X(J/ X/3
)
+ ...
(10.2)
r=O
- aw) Xa ~
Va /3
r=O
= aa'v a ) X(J/ :r:/3
we have
E = V(O) j
dr + L Vo' j XO'g dr + a
VO'/3 =0 if
r=O
(10.5)
Moreover, V must satisfy Laplace's equation:
\7'V = 0
(10.6)
This equation, evaluated at the origin, gives us
LVcrO' = 0
" 486
which, combined with (10.7), makes all three derivatives zero. The quadrupole coupling then vanishes. This situation arises, for example, with Na 23 in Na metal. The face-centered cubic crystal stnlclure puts each nucleus al a site of cubic symmetry. It is convenient to consider the quantities QO'(j defined by the equation 2
(10.9)
Qa/3 = j(3xcrx/3 - ocrpr )gdr
(10.10)
j x a x/3gdr = !(Qa/3 + j 8cr /3r gdr)
As we shall see, the introduction of the QO'p'S amounts to our subtracting from the left side of (10.10) a term that does not depend on the orienllltion of the nucleus. We have, then, for the quadrupole energy E(2),
;!
a f. {J
(10.8)
(cubic symmetry)
2
(10.3)
L VO'/3 j XO'x/3g dr... . (10.4) cr ,/3 Choosing the origin at the mass center of the nucleus. we have for the first tenn the electrostatic energy of the nucleus taken as a point charge. The second tenn involves the electrical dipole moment of the nucleus. It vanishes, since the center of mass and center of charge coincide. That they do coincide can be proved if the nuclear states possess a definite parity. All experimental evidence supports the contention that nuclei do have definite parity. Moreover, a nucleus in equilibrium experiences zero average electric field Va' It is interesting to note that even if the dipole moment were not zero, the tendency of a nucleus to be at a point of vanishing electric field would make the dipole tenn hard to see. In fact it was for just this reason that Purcell, Ramsey, and Smith [10.1] looked for signs of a possible nuclear electrical dipole moment in neutrons rather than in charged nuclei. The third tenn is the so-called electrical quadrupole tenn. We note at this point that one can always find principal axes of the potential V such that (!
Vu = Vyy = V.;z
In tenns of the QcrP's, we have
where Xo' (a=1,2,3) stands for x, y, or z, respectively. Defining
V."
dent part of the potential, and must therefore subtract the spherically symmetric parts.) If one has a nucleus at a site of cubic symmelry,
(10.7)
E(2) =
~L
V a/3
j xo'x/3(! dr
",~
=
~ L(Va/3Qa/3 + VO'/38a !3 j
2
r gdr)
(10.\1)
",~
Since V satisfies Laplace's equalion, the second tenn on the right of (10.11) vanishes, giving us
E(2) =
~ L: V"~Q"~
(10.12)
",~
Even if this tenn were not zero, we note that it would be independent of nuclear orientation. 2 2 If there is an electronic charge at the nucleus, we must usc Poisson's equation. Thcn
LV.... =
_'hrel,p(O)1
2
" 2 where 1,p(0)1 is the electronic probability density at the nucleus. The orientation independent term, A.E, of (IO.lt) becomes LJ.E
= ~ EV.... j
r20dT
=_4;e J,p(O)J2 j
r 2qdT
"
This LJ.E wi\.l...bC"different for two nuclei of the same c1l11rge but different charge distribution~'iliOi:opcs), or for two nuclei of the same mass and charge but different nuclear ~a.tes'(iSomers).In an electronic transition between an 8 and a p-slate, LJ.E will make a contribution that will in general be different for different isotopes or isomers. Effects also show up in nuclear transition [10.21. 487
To obtain a quantum mechanical expression for the quadrupole coupling, we simply replace the classical e by its quantum mechanical operator e(op), given by e(OP)(r) '"
1:: qAo e(r -
•
rAo)
,
(10.13)
where the sum runs over the nuclear particles, 1,2, ... k . .. N, of charge qAo' Since the neutrons have zero charge, and the protons a charge e, we can simply sum over the protons: e(Op)(r) '"
e
1::
S(r - rk)
By substituting (10.14) into the classical expression for QOlf3' we obtain the
1::
J(3:t a zfJ - 6a fJ r2 )S(r - rk)dT
protOll!l
1::
'" e
(10.17)
where' C is a constant, different for each set of the quantum numbers I and '1' In order to justify (10.17), we need to digress to discuss the Clebsch-Gordan coefficients, the so-called irreducible tensor operators TLM' and the WignerEckart theorem.
10.3 Clebsch-Gordan Coefficients, Irreducible Tensor Operators, and the Wigncr-Eckart Theorem
Q~;>:
Q~1) '" JO%et%f3 - Detpr 2 )e(oP)('r)dT '" e
(lm"IQ~;)IIm'") = C(lml~(loIp + Iplo ) - 'opI'IIm')
(10.14)
protons
quadrupole operator
These can be shown to obey the equation
(3ZakZ,Bk - Set,Br:).
(10.15)
protonll
We have, then, a quadrupole term for the Hamiltonian 1tQ' given by Q(oP) 'liQ = (51 '" L..- Vop .p
o,p
(10 16)
.
The expressions of (10.15) and (10.16) look exceedingly me~sy to handle because they involve all the nuclear panicles. They appear to reqUlr: us t~ m:at the nucleus as a many-panicle system, a complication we have aVOIded In diScussing the magnetic couplings. Actually a similar problem i.s involv~ in ~h magnetic dipole and electric quadrupole cases, but we have Simply aVOIded diScussion in the magnetic case. The quadrupole interaction represenled by (10.15) enables us to treat problems of much greater complexity than those we encounter in a discussion of resonance phenomena. When performing resonances, we are in general concerned only with Ihe ground stale of a nucleus, or perhaps with an excited state when the excited state is sufficiently long-lived. The eigenslates of the nucleus are characterized by the tolal angular momentum I of each state, 21 + I values of a component of angular momentum, and a set of other quantum num~rs '1, which we shall not bother to specify. Since we shall be concerned only With the spatial reorientation of the nucleus for a given nuclear energy state, we shall be concerned only with matrix elements diagonal in both I and '1' Thus we shall need only matrix elements of the quadrupole operator, such as
The Wigner-Eckart theorem is one of the most useful theorems in quantum mechanics. In order to state it, we must introduce the Clebsch-Gordan coefficients C{LJ'J;MMJ,MJ), and the irreducible tensor operators TLM' We shall first state the Wigner-Eckart theorem and then define the Clebsch-Gordan coefficienls. Next we shall discuss the irreducible tensor operators, and lasdy we shall indicate the derivation of the Wigner-Eckan theorem. We consider a sel.0f wave func.lions characterized by quantum numbers J and J' for the total angular momentum, M J or My for the z-eomponenl of angular momentum, and as many other quantum numbers '1 or 1'/ as are needed to specify the state. We are then concerned with calculating the matrix elements of the operators TLM' using these functions as the basis functions. The WignerEckart theorem states that all such matrix elements are related to the appropriate Clebsch-Gordan coefficients through a set of quantities (J'111 T L II J'1'/) that depend on J. J', '1. '1' and L but which are independent of M J' M J" and M. Stated mathematically, the Wigner-Eckart theorem is (JMJ"ITLMIJ'MJ") = C(J'LJ; MJ'M MJ)(J" II T L II 1'"')
.
(10.18)
Lei us now define the Clebsch-Gordan coefficienls. They are encounlered when one discusses the addition of two angular momenta to form a resultant. We therefore consider a system made up of two pans. Let us describe one part of the system by the quantum numbers L and A!, (Q describe the tOial angular momentum of that part and its z-component. Let us use the quantum numbers J' and MJI correspondingly for the second part of the system. For the system as a whole we introduce quantum numbers J and M J . We have, then. wave functions 'l/JLM and ,pJlM ,,10 describe the two parts, and tPJMJ for the whole system. The function tPJA:J can be expressed as a linear combination of product functions of Ihe two parts, since such products form a complete set:
tPJMJ '"
L
C(J'LJ;MJlMMJ),pJ'MJ,tPLM
(10.19)
J'MJ,;LM
(Im'1IQ::)IIm''1) 488
489
The coefficients C(J'LJ; M)'MM J ) are called the Clebsch·Gordtln coefficicms. Certain of their properties are very well known. Forexample, C(J' LJ; M)'M M J ) vanishes unless MJ "" M + M)'. A second property, often called the triangle rule, is that C(J'LJ;M;MMJ) vanishes unless J equals one of the values J' + L. J' + L - I, .. . IJ - Ll, a fact widely used in atomic physics. Let us now define the irreducible tensor operators Tu". Suppose we have a system whose angular momentum operators have components Jz, J". and J:. We define the raising and lowering operators J+ and J- as usual by the relations
J+ == Jz +iJlI
J- == Jz - iJ"
(10,20)
One can construct functions G of the operators of the system and examine the commutators such as [J+, G], [J- ,G], and [Jl' G]. It is often possible to define a family of 2L + I operators (L is an integer) labeled by an integer M(M ::: L, L - 1, _.. - L) which we shall term irreducible tensor operators T,.M, which obey the commutation rules
[J±. T LM ) = JL(L+ I) [J•. Tur)
M(M ± I)TLM ± 1
We shall wish to compute matrix elements of the TLM'S. We are familiar with the fact that it is possible to derive expressions for the matrix elements of angular momentum from the commutation rules among the components. It is possible to compute the matrix elements of the TLM'S by means of (10.21) in a similar manner. Let us illustrate. We have in mind a set of commuting operators J 2 , J., plus others, with eigenvalues J, MJI and '1. We use ." to stand for all other quantum numbers needed. We wish to compute matrix elements such as
(10.21)
(10.25) By means of the commutation rule
[Jz,Tu·,]::: MTUf
(10.26)
we have (10.27.) B",
.
Tlo::: J l
- (JMJ "ITLAI J.IJ'MJ ,.,,') , .
.
(10.22)
2
::: (MJ - M)')(JMJ1JITLMIJ'M)',,')
Another example of a Tl/Ir can be constructed for an atom with spin and orbital angular momentum operators sand " respectively, and total angular momentum J. Then we define the operators
1+:::/z+ilu
r:::lz-il u
One can then verify the operators T tM defined by 1 1 Til ::: - ..j2'+ TJO::: I: TI_l ::: ,jir
v
I
An example of such a set for L ::: 1 is -I + Til ::: ..j2J
.
(JMJ."I[Jz,TI.MlIJ'MJ,,,') ::: (JMJ"IJ.TutlJ'MJ ,,,')
= MTuf
(10.23)
(10.24)
obey (10.21). (Actually the operators of (10.24) form components of an irreducible tensor TIA'f with respect to the operators 1+, 1-, and Il as well as J+ J-, and Jz.) We may write the Tl/Ifs of (10.22) as TI/",(J), to signify that they are functions of the components Jz , J", and J. of J. The Tl/If'S of (10.24) are in a similar manner signified as T1J\f(l). It is helpful to have a more physic"l feeling for the definition of the operators TLAf by the commutation rules of (10.21). We realize that angular momentum operators can be used 10 generate rOiations, as discussed in Chapter 2. It is nOi surprising, therefore, that (10.21) can be shown to guarantee that TLM transforms under rotations of the coordinate axes into linear combinations TUf ', in exactly the same way that the spherical harmonics YL/lf trnnsform into linear combinations of YUI ' 's. This theorem is shown in Chapter 5 of Rose's excellent
Equation (10.27b) shows that
(JMJl1ITUI1J'M)'11')=0
unless
MJ-M),:::A{
(10.28)
I
book [10.3]. 490
In a similar way we may find conditions on the matrix elements of the other terms of (10.21). Thus (J MJ.I(J±. Tl.AII1J' Ml"') = JL(L + I) - M(M
± 1)(JMJ.ITl.Af ± i1J' M J ,"')
(10.29)
(J MJ"IJ±Tu"IJ'M J 'l1')
= (JMJ.IJ±IJ MJ 'f I.)(JMJ 'f I"ITLMIJ' Ml"") =
J J(J + I) -
(MJ 'f I)MJ(JMJ 'f I.ITLMIJ' Ml"')
(10.30)
By combining (10.29) and (10.30), we obtain the other recursion relations: 491
trix elements of Tr,M 's are related. The relationship is called the \Vig"er-Eckart
../J(J + I)
(MJ" I)M;{JMJ" hlITwIJ' M J,"')
- (J MJ'1ITLA1IJ' MJ'
=../L(L+1)
± 1'1')';JI(Jf + I)
theorem:
MJ'(MJ' ± 1)
M(M±I)(JMJ '/ITuf±dJ'MJ , " ) .
(10.37)
(10.31)
We note thal the only nonvanishing ICnTlS must satisfy (10.27b). However, if any one term in (10.31) satisfies this relation, all do. Equation (lO.27b) and (10.31) constitute a set of recursion relations relating matrix elements for TLM to one another and to those of TL/l1" 111cse equations lurn OullO be sufficient to enable one to solve for all TLM matrix elements for given J, )', lJ,1/ in ternlS of any onc matrix element. A further insight inlQ the significance of the recursion relations is shown by returning to the Clebsch-Gordan coefficients. In so doing, we shall sketch the proof of the Wigner-Eckart theorem. As is shown by Rose, the C's obey recursion relations identical to those of the TL/lf'S. We shall derive one: the selection rule on M. M;. and MJ'.
where the notation (JI'/II TL II J',./) stands for a quanlity that is a constant for a given J,£,J',I'/,I/ independent of M;. MJ'. and M. As we can see specifically from (10.22) and (10.24), for a given L and AI there may be a variety of functions, all of which are Tr,M'S. The Clebsch-Gordan coefficient is the same for all functions TUf that have the same Land M, but the constant (JI'/II T/, II J',/) will depend on what variable is used to construct the TLAI·S. To illustrate this point further, let us consider a particle with spin 8 and orbital angular momentum I and position r. The total angular momentum J is given by J::: 8+1
Consider the operator
J: ::= L: + J~
1.
(10.32)
where
(10.33) J~¢J'MJ' = MJ'¢J'M J, TIlen. using (10.19). consider the following matrix element of the operator J:: (tPLMtPJ' M J,' J:tJiJ MJ) = MJ(!/;LMtPJ' AlJ,' 1Ji; (HJ) =MJC(J'LJ;MJ'MMJ )
(10.34)
where we have let J: operate to the righl. But. writing J: as L:+J~ and operating on the functions to the left. we get (wur¢;'M J,' Jzwn.r) = (M + AJJ')C(J'LJ; M;,M M;)
(10.35)
By equating (10.34) and (10.35), we find (M + MJ' - MJ)CU ' LJ; MJ'M MJ) =
a
(10.36)
This equation is quite analogous to (1O.27b). provided we replace 1
(IMjIJIT/,flIIJ J\!!J"/)
By
~(z~ -%~) i ax 8z
I,
~(%~ -y~)
=
1
fJy
(10.39)
ax
We shall now list two T2M 's: one a function of the angular momentum J; the other, of the coordinate r. One can verify that the functions of Table 10.1, which we shall call Tu,r(J) and T 2M(r), indeed obey the commutation rules of (10.21) with respect to J+, J-, and J:. We have used the notation T2M (r) as shorthand for a TU,f constructed from the components z, y, and z of r. There is an obvious similarity between T2 Af(J) and T2M(r): Replacement of J+ by (z+iy), J- by (x-iy), Jz by z will conven TU1(J) into T2M(r). This similarity is a direct consequence of the similarity of the commutation relalions of components of J and r with J7.' J II , and J z : Table 10.1
by
T"
.
One can proceed in a similar manner to compute matrix elements of the raising and lowering operators, 10 get equations similar to (10.31). In fact the C(J'LJ;M;,/",fM;)'s obey recursion relalions identical to those of the (J M;I'/ITLAI IJ' M;,,/)'s. As a result, one can say that Ihe C's and the ma492
Bz
1 = 11
L:'PLM = MtPr,M
-
(10.38)
~ (y~ z~) I
J:oJi;M J = M;oJi;MJ
C(J'LJ;MJ,MMJ)
=
where
1'21
T" T'_I 1'2_2
J+'
-(J.;+ +J+;.) # ( 3 ) : ' _ ;') J.;- +;- J. J-'
493
[Jz.y] = iz
(10.40.)
[Jz • J y ] = iJ,
I
etc.
(1O.40b)
where (10.403) can be verified by means of (10.38) and (10.39). It is clear Ibat any function G(x. y. z) of x. Y. z, constructed from a function G(Jz , J y, J,) of J z , J1/' J, by direct substitution of x for J z , and SO on, will obey the same commutation rules with respect 10 J z • Jv' and Jz. Thus, if a function of J ZI JVI J, is known to be a TuJ, the same will be true of the function fonned by replacing J z • J 1I • J, by X. Y, z. respectively. The only caution we note in procedures such as this is that we must remember that the components of some operators do not commute among themselves; so that for example in T21 (J) we have the symmetrized product J+ J, + J,J+, not 2J+ J,. The method of direct replacement will work for olher variables as long as they obey commutation relations such as those of (10.40). For an excellent review anicle including tables of T/,M's of various Land M. sec [10.4]. Returning now to (10.37), let us consider two Tu",s, one a function of variables q and the other a function of variables p. Then (10.37) tells us thai 1
Iz =
, 1 (hili TL
II J',/)
II J'"')
(10.41)
Since the factor (Jf/ll TL(q) lJ J'11')/(J'111 TL(P) II J',./) is a conSlant (that is, independent of M, MJ. and MJ'), we see that we can compute all lhe matrix elements of TL}.f(q) of fixed J, J'. fl. and rl from knowledge of the constant and of the matrix clements (J M J '1ITLA1 (P)lJ' M Jl 11'). One word of caution is necessary. It may be thai (10.41) is nOI meaningful. since for some operators p. the matrix element (JMJ'1ITLi\,,(p)IJ'MJI '1') vanishes even though the matrix element (J MJ'1ITLM(Q)IJ' M J ,'1') does not. An example of such a case is when TLM(P) is made up of componenlS of J. Then all matrix elements in which J' .. J vanish. Of course (J 11 T2(J) II JI) vanishes 100. so that (10.41) becomes indetenninanl.
We now apply the Wigner-Eckarl theorem to evaluate the matrix elements of
oal3r~)
.
(10.42)
k(pl"Otons)
By recalling that I z • IIJI and I, are the operators of the total angular momentum of the nucleus 494
etc., for Iv and I,
,
(10.43)
where l z l.: and Szl.: are the x-components of the orbital and spin angular momenta of the kth nucleon; and by recalling that [tzkoYk)
=iZk
(SzkoYk]=O
,
etc. •
(10.44)
we see that [Iz.Yk)=izk
eiC.
,
(10.45)
The lenos 3xokXPk -ooprl are linear combinations of T 2M (rk)'S such as found in the right-hand column of Table 10.1. Equation (10.41) applies in a somewhat more general fonn not only to TLM'S but also 10 functions that are linear combinations of TLA's, all of the same L. Thus consider such a function F(P), which is a function of the operators p' F(p) = LClft",TU"(P) . (10.46) M
C(q) "
(10.47)
L"MTLA/(q) M
Then one can easily verify, using (10.41,46,47), that (JMJ"IC(q)/J'M J ."') = (JUJ"IF(P)IJ'Up ,i) (J" II TL(q)
II J',J')
x (J" II TL(P) II J''I')
(10.48)
We may apply this theorem to show thai (Im'1le
L
(3xokX,8k - oo,8r~)Jlm''1)
k(proloOS)
= (lrnql3
(lT"I<
(3XO'kX,8k -
k '
(l.T~
+ TpT.) 2
- '.pT'ITrn'q)C
(10.49)
where C is a constant,3 the same for all m. m', 0', and p. We can express C in tenns of the matrix element for which m = m' = I, 0' = f3 = z as follows:
10.4 Quadrupole Hamiltonian - Part 2
L
3z
Let us define a function C(q) of lhe operators q, using the same coefficients aM:
(J MJ11ITLM(q)IJ MJ"/)
Q~;>. Now Q~1) = e
L, lzk +
L (3,1- '·1 ITT")
= C(TT"13T; _ T'ITTq)
k(protons)
=CT(2T-l)
.
(10.50)
Since the quantum number '1 is llssumed to be associated wilh a variable that commlltes wilh 12 and I,. we can omit it in evaluating the right-hand side of (10.50). We shall also define a symbol eQ: 3 Do not confuse C with th~ symbot for the CJebsch-Gordan coefficients.
495
L:
oQ· (II.,I'
(3zl- rIIII")
(10.51)
k(protons)
have. by combining
Q is called the Quadrupole momellt of the nucleus. We (10.50) and (10.51),
C = 1(2l- I)
(10.52)
The fact thai we are concerned with matrix elements internal 10 one sel of quantum numbers I, 1/ enables us to use (10.49) and (10.52) to replace in the Hamihonian. All matrix elements diagonal in I and 11 are just what we should calculate by adding an effective quadrupolarcontribution 1-tQ to the Hamiltonian:
Q:;>
.
(10.53)
.,/J
II is interesting (0 note lhat of the nine components of Q~o;>. only one nuclear constant, eQ, is needed. The reason is as follows: The faci Ihal the nucleus is in a state of definite angular momentum is equivalent to the classical statement Ihal the charge has cylindrical symmeuy. Taking z as the symmetry axis, the energy change on reorientation depends. then, only on the difference between the charge distribution parallel and transverse to z: / z2(}dT
and
/ z2(}dT
.
This gives us the critical quantity /(z2 _ z2)(}dT
"'! /(2z 2 _ x 2 _ y2){ldT '" ! /(3z2 _ r 2)gdT
(10.54)
The last integral, we see, is the classical equivalent of our eQ. The effective quadrupole interaction of (10.53) applies for an arbitrary orientation of the rectangular coordinates a '" z. y, z. The tensor coupling to the symmetric (in z, y, z) tensor VO{j tan be simplified by choice of a set of principal axes relative to which Vo{j '" 0 for a oF {3. In terms of these axes. we have 1iQ '"
6I(;~- I) [Vu(3I; -
2 2 1 ) + VIIII (3l; - 1 )
+ Vu (3I"2 - l 2 )]
(10.55)
This expression can be rewritten, using Laplace's equation 1iQ '"
4I(;~- 1)[Vu (3l; -
Eo V
1 ) + (V~z - Vyy)(I; - I;)j 2
CfO
= U, to give (10.56)
Equation (10.56) shows that only two parameters are needed to characterize the derivatives of the potential; Vu and V:r:z - Vyy . IE is customary to define two 496
eq'" V" '1'", Vu - VIIy
'Q
'Q 1) L.. " V.pl,U.lp J 'liQ' 61(2l + lpl.) - 5.pl2 )
symbols, '1 and q. called the asymmetry parameter and the field gradient. by the equalions
(10.57)
V" The case of axial symmetry. often a good approximation. is handled by taking the axis to be the z-direction. giving '} '" O. Since we have seen that the raising and lowering operators often provide panicularly convenient selection rules. it is useful to write (10.53) in tenns of I+, I-. and I, for an arbitrary (that is. nonprincipal) set of axes. By defining
Vo
",Vu
V± 1
'"
V±2 '"
V,z
± iV,~
(10.58)
levu - VlIlI}±iVzlI
we find by straightforward algebraic manipulation that
'Q(22 -1iQ'" 41(21 _ I} Vo(3!., - I ) + V+1(l I, + I,I )
+ V_I (.P" I" + I"I+) + V+2(r)2 + V_2([+)2)
(10.59)
Equation (10.59) gives a form of the quadrupole coupling that is particularly useful when considering relaxation for which the principal axes are not fixed in space but rather are functions of time. An anempt to use principal axes would then be exceedingly cumbersome. We shall not attempt to describe nuclear relaxation by the quadrupolac coupling, although it is a very imponant mechanism in insulating crystals, often dominant at room temperature.
10.5 Examples at Strong and Weak Magnetic Fields In order to illustrate the use of the effective quadrupolar interaction. we shall make the simplifying assumption of a field with axial symmetry (or any other symmetry such that Vu '" Vyy for a set of principal axes). Let us then consider a magnetic field applied along the z'-axis where in general the z- and z'-axes differ. Then we have, for our Hamiltonian.
1i '" -'YnliHoI", +
4It22;~ I) (31; -
2 1 )
(10.60)
First we shall consider what happens when the quadrupole coupling is weak compared 10 the magnetic interaction. In Ihis case we can consider the spin quantized along the zl·axis. We proceed 10 U'Cat the quadrupolar coupling by 497
J<·ig.l0.Z. Axesz', z', and z
z' z 8
- 311
--r-
_1/ 2
-----------
'I, 111-% perturbation theory, Defining the x'-axis to lie in the plane containing z' and z, we have (see Fig. to.2) (10.61)
I:: '" I::, cos 8+lz ' sin 8 .
TH,
-r-
(b)
(al
Fig. JO.3. (a) Effect ora quadrupotfl coupling in first or
(a). The central line i$ unaffected by lhe quadrUI)()le coupling in first order
By substituting in (10.60), we have
e2 qQ 2 2 2· 2 1i'" -1nhHoI"+41(21_1)[31::.cos 8+3Iz .sm 8
-l
+ 3(l"lz ' + Iz,I::.) sin 8 cos 8 -12]
(10.62)
in this equation, since I::, is diagonal in first order and 1%. has vanishing diagonal elements, we have no contribution from tenns such as 1::,lz ' in first order. On the other hand, has diagonal elements, since it involves the product of offdiagonal elements. By expressing 1z ' '" !(1'+ + 1'-) and III '" (IJ2iXI'+ -
I;,
1'-), it is straightforward 10 show that the diagonal elements of ~. and 1~ are identical. We can therefore compute the diagonal matrix element
(mll;.lm) '" (mll;,lm) '" !(mI12 -1;.lm) = H[([ + I) - m') By collecting tenus, we find 2
e qQ Em'" -1n 1lHOm + 41(21- I)
(10.63)
[([ +
[')
.
i
(10.65)
Clearly 1 2 and I:: commute with 1i, giving the quanlum numbers 1 and m respectively. The energies are
1)[3m'-[([+I»)
(10.66)
!
(10.64)
,
i
4[;;;~ I) (3[; e2 qQ
(3COS228-1) [3m, -1(1 + I))
1» = T< (3[; - [') = 0
1/ =
E= '[(U
The effect of the quadrupole coupling is shown in Fig. 10.3 for the case of a spin I '" ~. It is helpful to note, since
I: [3m' -
the frequency of the other transitions. When a nucleus has a panicularly large quadrupole coupling Ihe chance is greal that even for well-annealed crystals one sees only lhe +! to transition. If we cany the percurbation to the next higher order even the to transition is shifted, the shift being of order (e2qQ)2hnhHo. A contrasting experimenul situation arises when the quadrupole coupling is larger than that to the magnetic field Ho. TIlen it is appropriate to consider the quadrupole coupling as a first approximation. We have for the Hamiltonian in the absence of an external field (still assuming axial symmetry),
m
lhat the quadnlpole coupling does nOI shift the center of gravity of the energy in first order. Moreover, lhe shifts of +m and -m are identical. With these points in mind we realize that the energy levels must look as shown. One interesting result is that for a half-integral spin the m '" levels are shifted the same amount and the t.ransition frequency between them is unaffected in first order by the quadrupole coupling. The! !O transitions are quite insensitive to effects such as crystalline strains which may tend to shift
A set of levels is shown for 1 '" in Fig. 10.4. We note that there is a degeneracy of ± m conesponding 10 the f::.ct that turning the nucleus end for end does not affecl the electrostatic energy. If an alternating magnetic field is applied with a nonvanishing component perpendic~ ular 10 the z·axis it produces nonvanishing matrix elements of ILlml '" 1. It can therefore produce resonant transitions between the quadrupole levels. It is
±t
±!
-!
498
±t ±l
Ftg.IO.4. Enflrgy levels of a quadrupole coupling when the Zeeman coupling is negligible
499
customary to speak then of "pure quadrupole resonance". although the transition is still induced by magnetic dipole coupling to Ihe allernating field. An imponant observation related to (10.66) is that when I is and so on (in general when 1= n +~, where n is an integer) Ihe energy levels are all doubly degenerate in Ihe absence of a magnetic field but that for integral spin the degeneracy may be completely removed, as with the m = 0 state. This result is an example of an imponant theorem. due to Kramers. and applies to both eleclTon and nuclear magnetic resonance. Kramers' theorem stales: For a system of angular momentum 1= n where n is 0, 1.2. and so on. the degeneracy of any state can never be completely lifted by electric fields. A corollary is that when a system is composed of an odd number of spin, particles. electric fields can never completely lift the degeneracy. The degeneracy is commonly called the Kramcrs' degeneracy. Proof of its existence actually depends on the propenies of the system under time reversal.
!' i,
+!.
10.6 Computation of Field Gradients We have seen that the quadrupole coupling depends on Ihe second derivatives Va ,8 of the potential, which reduce for the case of axes x. y. z. which are principal axes to V.u and V:n : - VylI . The potential V arises from external charges of either other nuclei or electrons. It is a straightforward matter of taking derivatives of the potential 10 show that a charge e at a point z. y. z produces a Vu at the origin of
Vu
=
e
r2 = x2
(3z 2 _ r 2 ) r
5
Vu·
=e
+ y2 + z2 (3 cos 2 8 - 1) r
,
Valli = -
(10.68.)
3
r
,
p = z.
y, z we have
, (3X.Xp ) - - -Ii • r a"
rJ
p
(lO.68b)
2
Equation (10.67) and (10.68) emphasize through the I/r 3 dependence that charges close to the nucleus have the most imponant effect. We may suppose that the electrons belonging to the atom containing the nucleus would make major contribUlions to Vu . Such is indeed Ihe case. However, if we have a closed shell. the electronic charge is spherically symmetric. and there is no quadrupole coupling (see. however. some funher remarks on closed shells below). The case of an incomplele shell is readily illustrated by an example of a single p-electron in an orbit zf(r). We wish, then. to compute the quadrupole operator for Ihis example. Since the electronic molion is mpid. we shall average the expression of (10.68) over the electronic orbit. This procedure is equivalent to saying that. of the total Hamiltonian describing bolh the electron orbil and the nuclear spin. we shall compute only matrix. elemetlls that are diagonal in the electron orbilal quantum numbers. and that we shall negieci the perturbation of the electron orbit by the nucleus. We have. then,
Vu
_
- -e
= -e
(10.67)
In terms of spherical coordinates we ha ve (see Fig. 10.5) 11"
In general for a or
: -,
/.,•• <3 CO,20 - I)
,.
'Pe
/
I~
3
tPcdTc
2 (3cos 28 - I) r 2 f2(r) cos 8 3 sin8r2drdfd8 r
C')
4x
(10.69)
where we have designated the electronic charge as -e and where. as usual, (117- 3) is the average of 1/,.3 for the p-orbit. We note that large-Z atoms. for which (l/r 3) is very large, will have large field gradients. This trend is shown by Table 10.2., which lists typical values of e 2 qQ of halogen nuclei in covalently bonded crystals. It is interesting to note that the values of e2 qQ of Table 10.2 put the f~ quency of pure quadrupole transitions for covalently bonded halogens at much higher frequencies Ihan that of Iheir Zeeman transitions "InBo for typical laboratory magnetic fields.
o Table 10.2. Typicl\l vl\luC$ or e2'lQ for hlllogen nuclei in covl\lently bonded crystals
Nucleus
/'
""- ,/ --------_':# ~
x 500
e 2qQ 11oo1llzi
y 80
SO. 2000
Q [10 24 cm 2 1 -7.97 X 10- 2 0.30 -0.59
Fig. 10.5. Location or a chiLrge ... in terma or the spherical coordinates r,',~
501
When the electronic wave function contains a mixture of s- and p-states (a "hybridized bond") the s-part conlributes nothing to the quadrupole coupling. A similar situation arises when a halogen atom is in a state corresponding to a mixture of a pure covalent bond (that is, a p-state) and an ionic bond (closed shell). The quadrupole coupling of the ionic bonding vanishes. One can therefore utilize quadrupole couplings to study bond hybridization, degree of covalency, double bonding, and so on. The fact that the closed shell electrons are very close to the nucleus makes it important to consider their distortion from spherical symmetry. For example, a charge e will produce fields that will disturb the closed shell electrons. This effect has been studied extensively by various workers. It leads 10 a correction to the gradient V:O: due to e alone. The actual field gradient V:: is in fact given by
(10.70) The quantity ,.(r) is called the Sternheimer antisliielding factor, after one of the workers who has made some of the mOst important contributions to understanding of the phenomenon (10.5). The fact that it is a funclion of the distance r from the charge e to the nucleus is emphasized by writing ,.(r). In general 1'(r) < I as long as e is well inside the closed shell charge distribution. Once r is well outside, i becomes independent of r. We shall denote this value by 1'00< Some theoretical values for I - 'Yoo are shown in Table 10.3. Table 10.3. Theoretical values of 1 - 700 Ion
J - 700
CI Cu+ Rb+
48 10 51
ea+
99
As we can see, the correction is enonnous, amplifying the direct effect V:o: by one or two orders of magnitude. The existence of the Stemheimer effect greatly complicates the determination of nuclear quadrupole moments. It is difficult to know how accurate the theoretical ,.'s are. However, examination of (10.68) reminds us of the magnetic dipole coupling between a nucleus and an electron spin. The radial and angular terms are the same as the A and B tenns of the dipolar coupling. Since the nuclear and electronic magnetic moments are known, it is therefore possible to use measured hyperfine couplings to get the average of (3 cos 2 8 _1)/r 3 . Since we are not using closed shells (the hyperfine coupling of a closed shell vanishes, since the electron spin is zero), the Sternheimer factor is only a small correction. This technique has been applied to the alomic beam experiments of halogens and is the basis of the most reliable experimental measurements of nuclear quadrupole moments. 502
11. Electron Spin Resonance
11.1 Introduction So far we have confined our attcntion to nuclear magnctic resonance, although many of the basic principles apply to clectron spin resonance. We have also considered questions concerning the electrons, such as the quenching of orbital angular momentum and the magnetic coupling of the nuclear spin to that of the electron. In this chapter we shall add a few more concepts that are imponant to the study of electron spin resonance l but which are not encountered in the study of nuclear resonance. Probably the major difference between electron and nuclear magnetic resonance is the fact that the nuclear properties such as spin, magnetic moment, and quadrupole moment are to a very high degree of approximation unaffected by the surroundings, whereas for electronic systems, the relatively much greater physical size and the much smaller energy to excited states make the system strongly dependent on the surroundings. An atom, when plnccd in a crystal, may have angulnr momentum, magnetic moment, and quadrupole moment valucs entirely different from those of its free atom. It is as though in nuclenr resonance we had to compute 'YII' I. and Q for each material in which the nucleus was to be studied. The fact that the state of an atom in a solid or liquid is very different from that when it is free means that we cannot predict the properties or even the existence of a resonance from the free atom electronic angular momentum and magnetic moment. For example, a sodium atom has zero orbital magnetic moment and angular momentum, but it has a spin of and a corresponding spin magnetic moment. The magnetic properties can be studied by the method of atomic beams. In sodium metal, the valence electrons fonn a conduction band, with substantial pairing of spins. However, there is a weak electronic spin magnetization whose spin resonance has been studied. In sodium chloride, the sodium gives up its outennost electron to complete the unfilled p-shell of the chlorines. The result is a zero spin magnetization and no electron spin resonance. Even if one has atoms whose bonding is covalent, as in molecular hydrogen, there is usually
i
I Seoe ~r~rencdl to -Electron Spin Il.eson"nce- listed in the Bibliography.
503
no net spin magnetiuuion because the electron spins pair off into a spin singlet. There are exceptions, of course, such as the oxygen molecule. As we remarked in connection wilh chemical shifts, the orbital angular momentum is often quenched, so that there is no first·onler orbital contribution to a resonance. We see Ihal mOSt insulators will nOt exhibit a resonance, unless one takes special pains to unpair the spins. Some atoms, such as those in the iron group or rare earths, have incomplete inner shells. Even when ionized, they still possess a net moment Thus neutral copper has a configuration (3d)104s. Cu++ has (3d)9, which is paramagnelic. In an ionic substance such as CuS0-t· 5H 20 (copper sulfate), the copper atoms are paramagnetic, and a resonance results. We may list several classes of substances or circumstances in which one may expect to find resonances, although in individual cases the general rules may break down: I. 2. 3. 4.
Materials containing atoms of the transition elements with incomplete inner shells; as, for example, the iron group or rare earths. Ordinary metals, the conduction electrons. Ferro-- and ferrimagnets. Imperfections in insulators, which may trap electrons or holes. For example, the F-center (electron trapped at the site of a missing halogen ion in an alkali halide) or donor and acceptor sites of semiconductors.
Tre'Hment of all these situations on a unified basis is SO hopelessly general that none of the interesling features emerge. The approximations important in one problem may nOt be at all justified in another. For example, if one is dealing with the resonance due to Cu++, one knows already a great deal about the electronic wave function, since it will be closely related to that of a free Cu++ ion. One can therefore start by considering states of a free copper ion. On the other hand, there is no equivalent to the "free ion" if one is dealing with an F-center. We could not, therefore, define a set of "free ion" states. What we shall do is list some of the more important interactions and then consider several examples that represent rather different physical situations but which involve the major phenomena. The principal teons in the electron Hamiltonian will consist of: I. 2.
3.
The electron kinetic energy. The electron potential energy. Often it is convenient to divide this into a "free ion" potential energy plus one due to the crystalline surroundings, the so-called crystalline potential. Such a decomposition makes sense provided there is such a thing as a "free ion", but, as remarked above, it would not have meaning for an F-center. The spin-orbit coupling. An electron moving in an electric field E expe· riences a coupling of the spin to the orbital motion 'HSO:
eh
?iSO' - -2 5 · (E x p) 2m c 2 S04
(11.1)
Often the electric field in an atom points nldially outward and is a function of r only, so that r
E(r) = -E(r) r
Then Ex p becomes (l/r)E(r)r x p '" (1IIr)E(r)L. This circumstance leads to the well-known fonn of the spin-orbit coupling, utilizing the spinorbit coupling constant ,),:
'Hso",,),L·S
4. 5. 6.
.
(11.2)
For free atoms that obey Russell-Saunders coupling, the spin-orbil coupling gives rise to the spliuing of states of given Land S and their classification according to the total angular moment J '" L + 5, L + 5 - I, ... , IL - 51. The coupling of the electron spin and orbital magnetic moments to an externally applied magnetic field. The magnetic coupling of the nuclear spin to the electronic spin and orbital moments. The coupling of the nuclear electrical quadrupole moment to the electronic charge.
Let us tum now to an example that will illustrate the role of some of the more important tenns. We begin in the next section with a discussion of the role of the crystalline fields and spin-orbit coupling. In the Section 11.3 we shall consider the coupling to the nuclear magnetic moment.
11.2 Example of Spin-Orbit Coupling and Crystalline Fields For our example we shall consider the case of an atom at the origin of a set of coordinates. possessing a single p-electron, acted on by four charges equidistant from the origin, two of the charges being positive, two negative, their magnitudes all being the same. The details of the arrangement (see Fig. 11.1) are seen to be identical to those thai we discussed earlier when we considered the phenomenon of chemical shifts. Neglecting nuclear coupling, we have for the Hamiltonian of the electron of charge q (q negative):
z y
+ +
•
lo'ig.ll.l. ArrangcmCllt or two positive lind two ncglltive charges, all equidistant rrom the origin
505
1 { =1- (
2m
<)' +Vo+V,+.\L·S+2/3H·S
p--A c
(11.3)
where A is the vector potential associated with the applied static magnelic field H, Vo is the potential of the "free alom", VI is me potential due to the fOUf charges, and 2f3H· S represents the coupling of the electron spin moment to the external field. We are here using fit the Bohr magnetoll, to express the electron magnetic moment. It is related to Ie. the electron gyromagnctic ratio, and J.f.e • the spin magnetic moment, by the equation Il.e = ~"Ye'lS:::: -2fJS
ie h =2fJ
I( 2m
•
p2
<)2 =---(p.A+A'p)+--A' q q2 p--A 2 c
2m
2mc
2me
( 11.5)
II is convenient to specify the vector p<>(cnlial as
A = ~Hx r
( 11.6)
which gives us
1( <)2 = 2m p2 -
2m p- ~A
qh
q2
2,2
2mc H . L + 8mc2 H (x
12
+y )
(11.1)
where, as usual, L = (I/i)r x 'V, and where the axes x' and y' are perpendicular to the field direclion Zl. (We distinguish between the field direction Zl and the crystalline axis z). The teml proportion to H2 gives the usual diamagnetism. It !Urns out to be unimportant, compared to the tenn H· L, in influencing the electron spin resonance. By utilizing the fact that f3 = eliflmc, we have. accordingly, as the Hamiltonian
p' 1{=-+,8H·L+Vo+VI +AL·S+2/1H·S . 2m
(11.8)
We shall consider that the principal energy tenns are the kinetic energy and the "free atom" potential Vo. We shall treat the remaining tenns by a perturbation method. For our example we shall think of the three degenerate I>-states xf('·), yf(1'), and zf(x), which are solutions of the free atom potential Vo. We shall assume that the coupling to other free atom states is relatively unimportant, so that the effect of the remaining tenns in the Hamiltonian can be found by considering only the submatrix of the Hamiltonian involving these three orbital states. For practical laboratory fields the tenns H· L and H· S are only about I cm- t whereas VI may be a substantial part of an electron volt (that is, 506
Alom
Coupling constant (em-II
B
10
C
28
F
211 440 1842
CI Br
( 11.4)
or
the negative sign representing the faCI thai the spin and moment are oppositely directed. Expanding the first tcnn on the right of (11.3) gives us -
Tabte I Lt. Spin-orbit coupling consbnts per electron for sevenol .toms
IOOcm- 1 to IO,OOOcm- I ). The spin-orbit coupling constants vary substantially. some typical values of the coupling per electron being given in Table 11.1. We see that, under some circumstances, VI will dominate; under other circumstances the spin-orbit coupling will be the major effect. The latter situation is. for example. typical of the rare earths. whereas the former is typical of the iron group. Let us first consider the case that VI is much larger than A. At first. we consider the effect of VI only. For the situation of Fig. 11.1, the effect of VI will be to lift the orbital degeneracy. The resultant energy levels, shown in Fig. 11.2, are all twofold degenerate because of the electron spin. We denote the wave functions as xf(r)u m • and so on. where the function Urn is a spin function. If there were no spin·orbit coupling. the spin would be quantized independently of the orbital state so that the urn's would be the usual eigenfunctions of S:/, where z' is the magnetic field direction.
yj(r)u.
zj(r)u", xj(r)u",
Fig.ll.l. Energy levels of the three p-sta~('$ under Lhe poLential of Fig. I l.l
Let us consider, then, the effect of the two remaining terms:
pH·L+AL·S .
(11.9)
We examine the sons of matrix elements these terms possess. "There are two SOrts: those that connect the same orbital state, and those that connect different orbital states. The former are clearly the more important, if they exisl, since the orbital splillings are so large. We have, then. matrix elements such as (11.10)
(11.11)
507
where ciT stands for an integral over spatial coordinates. and dTs. over spin variables. We see that both integrals of (11.10) and (11.11) involve
J
xf(,)L,xf(,')dT
(11.12) . X
Recalling our earlier discussion about [he quenching of orbital angular momentum, we realize Ihat the integral of (11.12) vanishes. Therefore the only nonva· nishing malrix elements of tenus pH· Land >.L· 5 connect states differing in the orbital energy. They have. therefore, no effect in first order. We have discussed this very problem in the absence of spin. noting that there was no first-order tenn pH· L. since the states xf(r) and so fonh correspond to no net circulation of the electron. A similar remark applies to Ihe spin-orbit coupling. The spin is coupled 10 states in which the electron has no preferenlial circulation. The average magnetic field due to orbital motion seen by Ihe spin vanishes. We know, however. from our discussion of chemical shifts that the teon pH· L will induce some orbital circulation. The spin will not. therefore, experience a suiclly zero field due to orbital motion. We can think of solving exactly for the wave function under the influence of the applied field and of computing the matrix elements of >.L· 5. using the exact wave function. As a practical matter we use penurba[ion theory to compute [he effect of pH· L on the wave function, keeping only the first tenn.' We have, then. for the modified wave function tPzm .pzm = xf(r)l1 m +
" (wm'IPH·Llxm) f() L.. w r 11 m' m' w=y,z Ez - E w
1::
(11.13)
Since pH· L does not depend on spin, m' = m. By writing the intera.ction pH, L in component fonn. we find .pzm =
J¢:m,>.L· 5t/Jzm ciT dTs
1
"" (wIL,lx) L.. E E pHqwf(r) [ x/(r)+ L.. w=y,z q=z,y,z z W
Um
(11.14)
We now use this corrected function to compute matrix elements of >.L ' 5 which involve Ihe ground orbital state. There are actually matrix elements to the excited state. which in second order can couple back down to the ground state. However, they do not involve the applied field H. By neglecting them, we are finding the field dependent coupling energy. (The second-order teons in spin-orbit coupling produce no splitting when the spin is
l)
2 We have two terms in the Ilamiitonian, {JH· Land >.L· S, neither of which gives a fil'!ll-order contribUlion. [n ~on(1 and higher ordel'!l, both terms perturb ll,e wave funClion. Bocause of the simi[arily belween lhe present problem and lhe ehemical shift, fil'!lt of all we trellllhe effoct of the term {JJI' L on the Wllve function. It may seem more sensible to take the much larger AL· S term first. As we shall see, however, our fllll\1 answer involves I'll l\PI>ro:o;iml\tion lhat is proportional to the product {J11' Land AL· S (it is the interplllY between the two energics th,,~ givcs the effe<:l). For this ptlrpOlle i~ is imm<\tcrial which interllction we usc to perturb the wave function.
E
= >.fJ
E
w=y,z q,q'=z,y,z
(m'IS,' Im)l(xIL,' Iw)(wIL,lx) + (xIL,lw)(wIL,' Ix)JH, E
z
E
w
•
(11.15)
For computing mauix elements internal to the ground orbital Slate, (11.15) is equivalent to our replacing the lerms plI· Land >.L· 5 by an effective teon in the Hamiltonian (1t'eff):
L
?t,rr ~
S,H,'),P
q,q'
~
L w
PL
[(xIL,lwXwILtlz) +(xILt1w)(wIL,lz)] E z Ell'
S,a,q'H,'
(11.16)
,.q'
Since the manix elements thai make up aqq' rransfonu under coordinate rotations like L q and Lq'. lhe aqq"s are components of a second-rank tensor. Examination of (11.16) shows that is a symmetric tensor (a qq , = aq'q). For our particular case we can compute the malrix elemenlS from the operators
L.
az - ay
~ ~(y~ ,~) I
L
Y
L,
~ ~(,~ I ax -x~) az
(11.17)
-
~ ~I (x~ ay y~) ax
By using these expressions, we find [hat
L.xf(r)
~
0 I
,
Lyzf(r) ~ 7,f(c) L,zf(')
~
(11.18)
I
,
-7yf(,-)
Thus the matrix elements of L z vanish, those of L y connect only to the state z/(1'), and those of L: connect only to yf('·). This gives us contribulions only from teons with q :::; q'. Thai is, the x-. y-, z-axcs arc principal axes of the leosor a'l'l" Specifically we have, using (11.16) and (11.18),
1t'err:::;211 (E :E SyHy + E >. E SzHz) (11.19) z z z y By combining this result with the Zeeman teon, 211H· 5, we obtain a spin Hamiltonian for the ground orbital state
508
509
H = {3(gzzH z S z + gyyHySy + gz: HzSz) where 9u = 2
9yy = 2 ( 1 - E
,
z
AEz)
We may employ the dyadic notation
9 10
(11.20)
g, defined by
= igzzi + i9 yyi + kgzzk
(11.22)
write the interaction as
'Ii=pH·
Ii
(11.23)
·5
in place of
'Ii = 2pH· 5
(11.24)
Comparison of (11.23) and (11.24) shows that the combined effect of the spin~ orbit coupling and orbital Zeeman energy is as though the real field H were replaced by an effective field Herr, given by
H
- H·
err-
2
9
-'H 9zz + 'H 9yy + kH 9zz Z 2 J y 2 z 2
-t
(11.25)
with the resonance given by
'Ii=2PH,rr. 5
.
(11.26)
Since gzz, gyy, and gzz are in general different, the effective field differs from the actual field in both magnitude and direction. If we denote by z" the direction of the effective field, it is clear that a coordinate transfonnation will put (11.26) into the fonn
'Ii = 2PH,rrS."
(11.27)
where Herr is the magnitude of Helf' The resonant frequency wo therefore satisfies the condition ~-,------:--:---:--;:
H;giz + H~9~y + H;g;z
= {3H
0~9iz + 0~9~y + o~g~z
(OI1t
(11.28)
where 01, 02 and 03 are the cosines of the angle between H and the x-, y-, and z-axes. Often one writes (11.28) as tlWo = g{3H
'H perl = AL· 5+ {3H· L
(11.29)
new
10') =
L /I
510
222222 gzzoJ + 9yy Cf 2 + 9z:03
(OI'li",wIO') =
Eo
(11.32)
E tl
L [(01.\£' 5In)("IPH· £10') + (OIPH· £1")("1.1£·510') n
+ (11.30)
(OIHperdn)(nl1t per d
By substituting from (11.31), we get
where the "g-factor" is defined by the equation g=
(11.31)
we find a matrix element (OI1t nc wIO') between 10) and 10'), as outlined in Appendix D, given by
flWO = 2{3He ff = {3
Equations (11.29) and (11.30) emphasize the fact that for a given orientation of H, the splitting of the spin states is directly proportional to the magnitude of H. Frequently one talks about the g-shift, a tenn that refers to the difference between 9 and the free spin value of 2. From (11.21) and (11.30), recognizing that both E z - E z and E y - E z are positive, we see that positive values of A make 9 less than or equal to 2, whereas negative A'S make 9 greater than or equal to 2. We associate positive A'S with atomic shells less than half-full, negative A'S with those more than half-full. Another tenninology is to remark that electron resonances give positive A and hole resonances give negative A'S. We shall return to this point in Sect. 11.4, where we shall find that a great deal of caution must be exercised in this simple interpretation as a general rule. The size of the g-shift clearly increases with the nuclear charge, as we noted in Table 11.1. Its magnitude also depends on the magnitude of the splitting to the excited states to which the orbital angular momentum couples. Using an energy of about 1.3eV (IO,OOOcm~l) and a A of lOOcm- 1 , we see 2 - 9 ~ 0.02, a readily observable effect. We note that the g-shift arises because of the interplay between the spinorbit and orbital Zeeman interactions. It is analogous to the chemical shift that arises from the interplay between nuclear spin-electron orbit coupling and the electron orbital Zeeman interaction. In both cases we say that the spin (electron or nuclear) experiences both the applied magnetic field and a sort of induced magnetic field. All such phenomena involving the interplay of two interactions can be viewed also as an application of a generalized form of second-order perturbation theory. This is in fact Ramsey's method for deriving the chemical shift fonnulas. We shall illustrate by computing the g-shift. The problem is treated in general in Appendix D. There it is shown that the perturbation effectively adds a tenn, 1t new , to the Hamiltonian, which has matrix elements between states 10) and 10/); for our example these elements have the same orbital pan xf(") blll may differ in spin function. Defining a perturbation leml 'H pert by
Eo
En
Eo
E/I
(OIPH· £In)(nIPH· £10') (01.\£·51")("1.1£· 510')] + . Eo E/I Eo - En (11.33) 511
The first two tenns on the right give the g-shift we have calculated. The last two tenns shift the twO spin states equally. They do oot, therefore, either produce a splitting of the doubly degenerate ground state or contribute to the g-shift. (If the spin were grealer than however, such a tenn could give a splitting of the ground spin state even when H:: O.) The last two teons are just what we should have had if either perturbation were present by itself. Our previous calculation of the g-shift did not give them because it treated the effect of one tenn of the perturbalion (PH· L) on the other (AL . 5). The method could be extended to find aU the tenns included in (11.33), but we see that direct application of (11.32) gives us a systematic method of getting all tenns. On the other hand, the physical principles of the first calculation are somewhat more apparent. In the example we have discussed so far, we have considered the crystalline potential VI to be much larger than the spin-orbit coupling constant A. As a result, the orbital angular momentum is largely quenched, and the g-value is very close to the spin-only value of 2. This situation corresponds to the iron group atoms as well as to many electron and hole centers. We tum now to the opposite case, one of strong spin-orbit coupling and relatively much weaker crystalline fields, as encountered in the case of rare earth atoms. If the spin-orbit coupling is dominant, the situation is in first approximation similar to that of a free atom. In fact the Hamiltonian
I.L· ~
4,
p' 'Ii = - + Vo +AL ·5+ pH·L +2fJH'5+ VI
(11.34) 2m is identical to that of a free atom except for the tenn VI. As a first approximation we consider just the effect of the tenn >.L· 5 on the state formed from spin functions and the three p-states xf(r), yJ(r), and ZJ(I·). The sum of the angular momenta Land S is the total angular momentum J:
J::L+5
.
(11.35)
By squaring, J, we have
~L·5= ~(J' -
2 with eigenvalues
Esc
~
L' - 5')
L(L + I) - 5(S + I»)
;:;;
L
Clm ,.Il'jm
(11.39)
I,m
where the YIIlI'S are spherical hannonics and the Clm'S are conslants. If the potential is due 10 the charges of Fig. 11.1, it vanishes on the z·axis and changes sign if we replace z by y and y by -z (a coordinate rotation). It is a maximum on Ihe x-axis and a minimum on the y-axis for a given distance from the origin. The lowest I in the series of (11.39) is clearly I:: 2. Of the five 1:: 2 functions, .:z:y, xz, yz, 3z 2 - r 2, .:z:2 - y2, only the last is needed. We have, therefore, as an approximation, insofar as teons for 1> 2 are not required, (11.40) where A is a constant. (We shall see shortly that no higher terms are needed for an exact treatment.) We have, then, to consider the effect of VI on the states of Fig. 11.3. Two sorts of matrix elements will be important: those entirely within a given J, such as (JM; IVtlJMJ), and those connecting the different J states. The fonner will be the more important because they connect degenerate states. We can compute the matrix elements internal to a given J by means of the Wigner-Eckan theorem, for we notice that, with respect to L z , L y , and L: (hence with respect to J z , JIJ , and J:), F I is a linear combination of Tu,,'s. That is, the commutation relations of VI with J:, JT. ± iJy show VJ 10 be a linear combination of T 2M 'so (In fact Vt is proportional to T22 + T2-2.) Thus we have
(11.36)
(JMJlVtlJM~)
=A(JMJlz' -
y'IJM'/)
:: CJ(JMJIJ; - J~IJMJ) (11.37)
Since we are concerned with states all of which are characterized by an orbital quantum number L and spin quantum number 5, the possible values of J are £+5, £+5-1, ... IL- 51. For our example, L:: 1, 5::::t!, so that J:: ~ or The J :: ~ and states are therefore split apart by an energy spacing
!
LIE = ~~
FI
given by
£So = 2:[J(J + 1) -
!.
We consider next the effect of Fl' Here il becomes convenient to assume a specific fonn. Assuming that the potential arises from charges external to the atom, Ihe potential in the region of the atom can be expressed as a sum of the fonn
(11.41)
This is equivalent to our replacing VI by the operator 1ft.
Hl :: CJ(J; - J;)
(11.42)
as long as we only compute matrix elements diagonal in J. As an allernative, had the potential VI been B(3.: 2 - r 2), we should have used
(11.38)
(11.43)
(More generally, the stale J is AJ above the state J - 1.) The energy levels are shown in Fig. 11.3 for a positive A.
This case of what is called the "axial field" is frequently encountered. We describe the calculation of CJ and CJ below.
512
513
We have still to consider the effect of Ihe magnetic field terms. Again lei us consider only the malrix elements dia{:onal in J. This means that we wish to have
(JMJIPH· L +2pH· SIJMj) = pH· (JMJIL +2SIJMj)
(J MJIL + 2SIJMj) = 9J(J M J IJIJMj)
(11.46)
Equation (11.45) is equivalenl (as far as malrix elements diagonal in J are concerned) to our having a term 'Hz replacing lhe two magnetic terms, where 'Hz is given by (11.47)
.
By combining (11.42) and (11.47) with (11.36), we obtain an effective spin Hamiltonian 1te fr' which describes our problem accurately within a given J: 'Hefr = C;(J; - J;)
~/2 =
~
(11.49)
The tWO tenns on the right of (11.48) and (11.49) lift the (2J + I)-fold degeneracy of each J Slate. We shall not discuss the details of handling this problem other than 10 remark Ihal clearly it is formally equivalent to the solution of the problem of a nucleus posscssing a quadrupole moment actcd on by an electric field gradient and a static magnctic field. So far we have not computed the constants C; or Cj. We lum now 10 Ihat task, illustrating the method by computing C We have. using (11.43)
(11.50)
By choosing /If; = Mj = J, we have
B(JJI3z' - r'IJJ) = Cj[3J' - J(J + 1)1
c'J =
J(2JB _I)(JJ I 3,' -r 'I JJ)
Now for J = 514
4, all matrix elements of 3J; -
(11.50.) (11.51)
J2 vanish (analogous to the faci
I 2 + y2)f2(r)(3z 2 - r 2)dr 2"(x
(11.53)
2B--r' 15
(11.53,)
!.
A(LML! z 2 - y2JLM~) = CL(LMLlL; - L~ILM~) 2 B(LMLl3z - r 2 JLM}) = GL(LMd3L; - L2ILML)
(11.54)
These matrix elements are equivalent to those we should have had by replacing VI by an equivalent Hamiltonian 1t1:
1t1 = GL(L; - L;) or 1t) = C'L(3L; - L 2)
(11.55)
In terms of 1tI, the effective Hamiltonian 1te rr. which will give all matrix elements fonned from the three p-states xf(r), yf(r), and zf(r), is
1te rr=>.L'S+C L(L;-L;)+PH.(25+L)
J.
B(J M;13z 2 - r21JMj) = Cj(JMJ I3J; - J2lJMj)
BJ
'3
i
or, for the axial field,
1tcfr=Cj(3J;-J 2)+9;pH.J
( 11.52)
where r 2 is the average value of r 2 for the p-states. The ~p.in. Hamiltonians of (11.48) and (11.49) do nOI include any malrix elements JO,"lng states of J = wilh those of J = If we wish to include such effects, we can actually apply the Wigner-Eckart theorem in an ailemative manner. All matrix elements of VI are between states of L = I. Thai is, we are concerned. only with (LMr,lVdLMJ). However, the commutators of L z , L y • and ~, with VI show that VI is a linear combination of T 2M 's. Therefore all matnx elements are of the form
(11.48)
+ g;pH· J
.
Th~ angular ~nions of the integral can be carried through, using spherical coordinates, to gIVe '""3/2::
J(J + 1) + S(S + 1) - L(L + I) g; = 1 + 2J(J + I)
~
where we have denoted Ihe spin function S~:: +~ by ul/2' Therefore we find
(11.45)
where 9; is a conSlanl for a given J, independent of M; or M J. We recognize this problem to be the same as that of compuling Ihe Zeeman effect of free atoms, and the constant 9; is therefore the familiar Lande 9-factor:
"/iz=gJPH·J
1
levels). For J :: 3, we
IJJ) = J2(%+'Y)
(11.44)
We can apply the Wigner-Eckart theorem to the malrix element, which, with respect 10 J, is a linear combination of TtM's. We can therefore write
i
thai a quadrupole coupling cannot split a pair of spin have
or
1teff :: >.(L· S) + Ci,<3L; - £2) + pH· (25 + L) .
(11.56)
Equation (11.56) reduces 10 (11.48) or (11.49) for matrix elements diagonal in
J. In the absence of an external magnelic field, the energy levels of Ihe Hamiltonian of (11.56) would remain at leasl doubly degenerate, according to Kramer's theore~, ~Iated i~ Section 10.5 since £ I, S It IS Interesung to note that the derivation of (11.45)
=
(JMJIL + 2SIJ Mj) = YJ(JMJIJIJMj)
=!.
(11.45)
using the Wigner-Eckart theorem. is quite analogous to the stalement about the 515
nuclear magnetic moment i-L and spin I (11.57)
which, stated more precisely, is (IMI"IIM') = 1t.(IMIIIIM')
(11.58)
We expressed the potential of Fig. 11.1 by means of only one tenn in the expansion of the potential, that with I = 2. We might have supposed also that tenns 1 = 4, 6, and so on, would have been needed. (The odd 1's are not needed for this example, owing to the inversion symmetry of the charges.) If we had included an 1=4 tenn, we should have then needed to compute matrix elements such as
(LMLlrIYtmILM~)
(11.59)
with 1 = 4. However, with respect 10 L1;' L y , and L z , rlYtmis a Tim' Therefore we can apply the Wigner-Eckart theorem 10 evaluate it. Recognizing that such an integral is closely related to the combination of angular momentum (by means of the Clebsch-Gordan coefficients), we note that it will vanish unless L and I can couple to fonn an angular momentum L (the triangle rule). For the case L = 1, I = 4, we can see that L and I could combine 10 give angular momenta of 5, 4, or 3, so that the integral must vanish. In fact, for L = I, only I = 2 gives nonvanishing matrix elements. We need not, therefore, bother with 1 = 4, 6, and so on, in the expansion of the potential.
connecting states of different crystalline energy such as x f(r) and y f(7'). We shall neglect the elements that are off-diagonal. We have, then, only matrix elements such as
(~msmJIHJSrlxmsml)
=
JtP~,u;"sxf(r)11.rSr , ,
x tPm,u m ' xf(r)drdrsdrJ
where ms and mr stand for eigenvalues of Sz and J z ; tPml and ums are nuclear and electron spin functions; and where dr and drs stand for integration, respectively, over electron spatial and spin coordinates, and drl over nuclear spin coordinate. It is, as usual, convenient to leave the specification of the quantization of both electron and nuclear spins until later, since the appropriate quantum states will depend on other parts of the Hamiltonian. We therefore omit the electron and nuclear spin functions and integrations, computing only J xf(7')Hrsrxf(7')dr. This integral leaves the nuclear and electron spin coordinates as operators. We shall therefore denote it by 11.rs: HJS =
J
xf(r)11.J Srxf(")dr
(11.63)
By substituting (11.61), we obtain HJS =iein1l2
J
,.13 [3(I ·2r)(s.r) r
~I'slx2f2(7')dr
(11.65)
We have not as yet considered the magnetic coupling of the electron to the nearby nuclei. The basic fonn of the interaction has been discussed in Chapter 4. We distinguish between s-states and non-s-states. s-states : 81f
,2
Non-s-states : li2 H = iein [3(l.r)(5.r) -1.5] rSr r3 r2
(11.60)
where, as usual, (1/1- 3 ) denotes the value of I/r 3 averaged over the state xf(r). If instead of a p-state we had an s-state. or more generally a wave function If; containing some s-state, we could compute a corresponding tenn Hrs arising from the {j-function coupling: HJS =
J
1f;*(r)H rsr 1/;(r)dr = 8 1f ")'<)")'011.2 11/;(0)1 2 I· S 3
(11.66)
The most general interaction is. of course, the sum of the couplings to the interactions of (11.60) and (J 1.61):
(11.61)
The effect of the hyperfine coupling can be illustrated by considering the example of the preceding section in which the orbital angular momentum was quenched (corresponding to a crystalline potential VI much larger than the spinorbit coupling A). In the Hamiltonian matrix there will then be elements of the hyperfine coupling diagonal in the electron orbital energy as well as those 516
(11.64)
The tenns such as JzxSyY will contribute nothing, since for them the integrand is an odd function of x or y. The other tenns can be expressed as the product of angular and radial integrals, giving
11.3 Hyperfine Structure
Hrs r = 3ieinll I· 5{j(r)
(11.62)
hrs
=
Jh
b(I')1 2 {8;
i<)inrt2I.5~(r)
+ 101"h'[3(I'T)(S'T)] -I'S}dT rJ r2
(11.67)
It will be linear in the spin variables Jz , Jy , I z and Sz, Sy, and Sz, being of the
general fonn 517
11.IS:
L:
(11.68)
ACiOlSoJol
Ci,Ci':;e,y,z but with the A's symmetric (that is, A Cia , = Aa'a)' One can therefore always find the principal axes such that A aa , is diagonal, with values Act and the hyperfine coupling given as (11.69)
We see that I z commutes with (11.73). We therefore take the states to be eigenfunctions of I z with eigenvalues mI. The first-order energy is therefore E.= fJguHmS - 'YnhHmI + AzmSmI
Since I 2 , 52, and I z and 5 z all commute with the Hamiltonian of (11.73) we can take the eigenfunctions to be a product of a nuclear spin and electron spin function: (11.75)
where the dyadic A is given by
A = iAzi + jAyj + kAzk
(11.70)
If we did not have quenched orbital angular momentum, it would be necessary to include the coupling of the nuclear moment to the magnetic field arising from the orbital motion of the electron. We should also need to choose a new set of basic electronic states to obtain the "spin" Hamiltonian, such as, for example, states that are eigenfunctions of J2. We can combine (11.69) with (11.23) to obtain the spin Hamiltonian that includes both nucleus and electron for a case of quenched orbital angular momentum:
'H=j3H·
9 ,S-"YnnH·[+[·
-
A·S
.
(11.71)
(If the nucleus experiences a quadrupolar coupling, a term 'HQ should be added.)
We shall examine the sons of effects that the couplings of (11.71) produce by a simple example. We nOle first that there is no reason for the principal
9
A
axes of and to coincide, although they do in fact for many simple cases. (Experimental situations have been reponed in which they differ). We shall assume that they do, for our example. We begin, moreover, with the assumption that H is parallel to one of the principal axes, the z-axis, so that (11.72) We cannot solve this Hamiltonian in closed form without making some approximations. We shall assume that the electron spin Zeeman energy J3H gu is much bigger than the hyperfine coupling energies A;e, A y , and A z . This approximation is frequently good whcn one has strong magnetic fields (2J3H = 10 10 Hz for H = 3300Gauss, whereas A is often 109 Hz or less). If we take the electron Zeeman term to be large, we see that 'H. commutes with Sz to a good approximation. We takc the eigenfunctions to be eigenfunctions of Sz, with eigenvalue ms. The terms A;eSzI;e and A y5 y I y then have no matrix elements that are diagonal in ms. We drop them from the Hamiltonian in first order. On the other hand, AzSzIz is diagonal in ms and must be kept. This gives us an approximate Hamiltonian: (11.73)
5'8
(11.74)
The possible transitions produced by an alternating field are found by considering the matrix elements of the magnetic operator 'HmO): 11.m (t)
=
("Yeh5z - "YnllIz)Hz cos
wt
(11.76)
between states such as those of (11.75). We find in this way that the 5 z part of 11.m(t) connects states with ams = ± I, amI = 0, whereas the I z ponion connects ams = 0, amI = ± I. We can consider these respectively to represent electron resonance and nuclear resonance. The transitions are allowed only if w satisfies the conservation of energy, which, by using (11.74) and the selection rules, gives for electron resonance We
=
gzzfJHo + AzmI h
and for nuclear resonance Azms W n = "YnHO + - h -
(11.77)
(11.78)
The effect of the hyperfine coupling on the eleclron resonance is seen to be equivalent to the addition of an extra magnetic field proponional to the z-component of the nuclear spin. Since the nucleus can take up only quantized orientation, the electron resonance is split into 21 + I (equally spaced) lines. If the nuclei have no preferential orientation, the lines corresponding to various values of mI occur with equal probability, and the resonance pattern looks like that of Fig. 11.4. If one looks at the nuclear resonance, the frequencies are given by (11.78). To interpret the expression, we must know whcthcr the nuclear Zeeman energy inhHo is larger or smaller than the hypcrfine coupling AzmS' since (11.78) gives both positive and negative frequencies. In the former case (Fig. 11.5) the resonance is split into 25 + I lines spaced Azlh apart and centercd on the angular frequency inHO. If the nuclear Zeeman energy is smnllcr than Azms, the two values of ms, ms and -ms, give rise to two lines occurring at !'ll:. 11.4. Absorption \'ersus frequency for the
electron resonance for the ease of a nucleus of spin .1/. The lines lire, in first approximation, el'lce<1
519
r--~-j
_I
l,.Ho
I-----;:;~
J<'ig.lI.5. NudcllT ~nanc' when the IlUclear Zeeman energy islllrger than drawn on the llSSumption of an eleetron spin of
IA. msl
!
The Hamiltonian becomes
I- 2,.H,-j
_I
A.
I
1t = 2{JH:,5,' -inuR:,I" + A z l z 5 z + A y I y S lI + A:I,S: ~
2A' Fig. 11.6. Nuelcar resonance for all electronic spin of when the hypcrfine coupling is larger than lht nuclear Zeeman coupling
t
(ll.80)
We shall continue to assume 2{JHo ;» A z • A y • and A, so that, to a good approximation, the Hamiltonian commutes with S". We now seek to find those other parts of the Hamiltonian that will be diagonal in a representation in which 5:, is diagonal. To do this, we express the spin components Iz.Iy • I" and 5 z , Sy, and S: in terms of the primed axes. (Actually it would be sufficient to express 5 z • 51/ and S, in terms of 5 z " 5y" and 5". However. the transformation of the nuclear axes enables us to see more readily that the nuclear quantization direction is not parallel to that of the electron). Noting that y = y'. we have
Sz: = Sz:1 cos 8+S" sin 8 w""
A.lmsl ± H h "'In 0
{I 1.79)
i.
For an electron with spin the result is shown in Fig. 11.6. An examination of Figs. 11.4. 11.5. and 11.6 shows that the electron resonance will enable one 10 measure the hyperfine coupling tensor and the nuclear spin, but will nOI by itself enable one to measure the nuclear moment. When combined with the results of a nuclear resonance, even the nuclear magnetic moment can be found. This latter feature is of particular importance in the slUdy of color centers because it enables one to identify the nuclear species that gives rise to the hyperfine spliuing, since the nuclear i'S are known from other experiments. The nuclear resonance, on the other hand. enables one to measure the spin of the electron system. We have so far restricted ourselves to an orientation of the static field along one of the principal axes of the g- and hyperfine tensors. Interesting new effects arise when the field lies along other directions: The axis of quantization of the nucleus becomes different from that of the eleclrOn and in fact depends on whether the electron is oriented parallel or antiparallel to the field. For the sake of simplicity we shall illustrate these effects for the case where the electronic g-factor is isotropic. We shall take the field direction i to lie in the x-z plane. where x, y, and z are the principal axes of the hyperfine coupling. The orientation is illustrated in Fig. 11.7.
Sy = Sy'
5 z = 5,. cos 8 - 5z:' sin 8 I z = Iz:' cos 8 + I" sin 8 I y = Iy' I, = I,. cos 8 - I z ' sin 8
(11.81)
By substituting these expressions into (11.80). we find
1t = 2{JHo5" - inhHoI,. + I z .5z ,(A z cos2 8 + A, sin2 8) + Iy'5y'AI/ + I,.5,.(A z sin 2 8 + A, cos2 8) + (1z:.5,. + I,.5z ' )(A z - A,) sin 8 cos 8
To first order. those terms in lhe hyperfine coupling involving 5 z • or 5y' possess zero diagonal elements in the 5" scheme of quantization and therefore may be omitted. In second order. they will contribute energy shifts of the order of A 2/2{JHo. where by A 2 we mean the square of a matrix element of order A z • A y , or A,. and the 2{JHo comes in because the "excited state" differs in electron spin orientation in the static field. The reduced Hamiltonian 'Hred' which results from dropping all lerms involving Sz or 5y • is then 'Hre<J = 2{JHoS" - 1'nliH01" + {Iz,(A z sin 2 8 + A z cos 2 8)
+ I'l'.(A z - A,) sin 8 cos 8j5z'
,
(11.82)
(11.83)
Of course 5 z' commutes with 'Hred' However, I" does not As far as the nucleus is concerned. 'Hrcd corresponds to a nucleus coupled to a magnetic field with components Hz •• H y" and H" of }f
H'l" = -_
----x'
520
x
Jo'ig.lI.7. Orientation of axes z, :, principal l'xes of lhe hyperfinc tensor A, relative lo lhe direction of the stalic field "
1
-~ (A,
1'n,l
. - A z ) Sin 8 cos 85,. (11.84)
H y ' =0
2 2 (AzSin 8+A:cOS H z '-H -O~ 1'n/l
8)s
:'
521
These expressions involve the operator 5 z" by which we mean, of course, that the effective field depends on the electron quantum state. By denoting the eigenvalue of 52' by rns. we have
Hz' (rns ) "" 1 (
A: - A z ) sin ~"
1
,
8 8) cos
1-'lg.It.8. The axes (:1:',::2) are rolaloo an angle 112 uis ~Ialive lo lhe aXelI (ZI, zd
q, aboul the 111
:<
ms
HII(ms) = 0
(11.85)
Hz,(ms> = Ho - (
All: sin2 (J + A z cos2 8) ~"
h
ms
The direction of quantization of the nucleus is clearly along the resultant of the effective field of (11.85) and not along the static field. The direction differs for the different ms values. The magniwde of the nuclear energy separation
~--------'I
L1Enuc1ear is as usual
Ll.Enuclear = inhHen{ms}
(11.86)
where Hcrr(ms) is the magnitude of the fields of (11.85) and depends on ms. The total energy of electron and nucleus is therefore
v+ '" COS(4)/2)u+ + sin(4)/2)u_
v_ '" - sin(#2)u+ + cos(#2)u_
( 11.89)
(11.87)
We may notc another interesting side effect of the lilt in the nuclear quantization axis as one varies ms. The nuclear functions tlmJ for one value of ms are different from those for anocher value of ms. owing to the change in the direction of nuclear quantizalion. We write them as umJ(ms) to emphasize this point. We can express the 2I + I functions tlm,(mS + I) in tenns of the 2I + I functions of um/ms) by the relalionship um/(ms +
1):<
L:amsm/m~um~(ms)
(11.88)
m;
where the a's are constants. We have, then, the possibility of transitions ms, mt to mS + I, m where m mt (since amsmJm~ is in general not zero). Such a transition represents a simultaneous nuclear and electron spin transition. Actually it does not really represent a nuclear orientation change. Equation (11.88) simply expresses the fact that a nucleus of given mJ relative to Hcrr(mS} must (since its spatial orientation is fixed during the electron flip) go to a mixture of rnr's when the electron orientation changes, the m/'s referring as they do to a different quantization direction. We can see this result explicitly by a simple example for a nucleus with spin Consider two different quantiz:ltion directions ZI and Z2 that make an angle q, with respect to each other; see Fig. 11.8. The Z2-axis is rotated relative 10 the zl-axis by an angle 4J (in the right-handed sense) about the YI :< Y2 axis. Let u+ and u_ be the eigenfunctions of I~, and v+ and v_ be the eigenfunctions of I z,. Then the functions v and It are related through the Clebsch-Gordon coefficients
r
!.
[J 1.11: 522
r"
[Equation (11.89) can also be derived using the methods of Chap. 2. A rotation of 4> about the y-axis can be treated by finding tlk solution of (2.73) for HI along the negative y-axis with -yH1t w '" ;.] Then, if the direction of the effective field acting on the nucleus when the electron spin points up is taken as the %1 direction. and the direction when the electron spin is down is called the Z2 direction, and if a and f3 are the electron up and down functions (the eigenfunctions of 5::,). the complete eigenfunctions of the Hamiltonian are the four functions ImsmJ) given by
1++) "'
I-
+) '"
1- -)
f3v+ '" (3u+ cos(q,!2) + (3u- sin(;I2)
(11.90)
= {3v- = -{3u+ ,in(¢/2) + {3u_ co,(¢/2)
[Explicit derivation of (11.90) plus determinalion of the directions Zl, Z2. and the angle 4> are given in Sect. 11.4.] Clearly the operator Sr.' will have matrix elements which join either C1'U_ or OU+ to both fJv+ and {iv-. So far we have treated the electron as being quantized along the direction of the static field Ho. There are two other refinemenls worth mentioning. The first is concemed with anisotropy in the electron g-faclor. If the electron g-factor is anisotropic, the electron will actually be quantized along the direction of the effective field acting on ii, H g • given by
523
(11.91)
This refinement will slightly modify the treatment of (I 1.81-84). The other refinement worth mentioning is to include the fact that the nucleus exerts a magnetic field on the electron. II therefore should playa role in the clec(ron spin quantization. Thus, if the nucleus is flipped, the electron quantization direction changes slighdy. This change is ordinarily small and can frequently be neglected. However, there are some situations in which it is important. Waugh and Stichter [11.2] have called this slight change in electron spin quantization "wobble," They point out thai it gives the electron a degree of freedom much less costly in energy than flipping between the stales 0' and p. They point out that it may playa role in nuclear relaxation by paramagnetic ions at very low temperatures since wobble can be excited by the small thermal energies, but electron spin flips between states 0' and (J cannot.
11.4 Electron Spin Echoes The use of pulse techniques, in particular of spin echoes and stimulated echoes, has become important in electron spin resonance. We turn now to some special features that arise for this case which are not encountered for spin echoes in nuclear magnetic resonance. As we have seen, if we have two coupled nuclei belonging to two different species (e.g. 13C and IH), the spin-spin splitting of one species produced by the other does not affect the size of the echo. Indeed, it acts just like a static magnetic field inhomogeneity. Therefore, although it affects the free induction decay, it is refocused perfectly at time 21" by application of the second pulse of the echo sequence. In order to observe the spin-spin coupling on an echo, one must do a double resonance experiment (Sects.7.19-21). For an electron (spin 5) coupled to a nucleus (spin I) the situation is different owing to the fact, which we have just discussed, that in many cases the nuclear quantization direction depends on the electron spin orientation. As we shall see, the consequence of this coupling is that the amplitude of the electron spin echo is modulated as a function of the time 1" between pulses by the 1- S spin-spin coupling. The modulation frequencies are the splitting frequencies of the nuclear magnetic resonance. These are the frequencies one observes in ENDOR. Therefore, the electron echo envelope in effect reveals the ENDOR frequencies. It is this special feature which has made spin echoes such an important technique for electron spin resonance. This phenomenon was first reported by Rowan et al. [11.3] and subsequently elucidated in greater generality by Mims [11.4 1. Before providing a formal description, let us examine physically what is happening. In particular, why is the electron echo case different from the nu-
clear echo case? For a pair of coupled nuclei of spins I and S respectively, the quantization of one nucleus is to an excellent approximation independent of the orientation of the other nucleus. Consider, then, application of a 1r/2 ... r ... 1r puls~ sequence to the 5-spins. If the I-spin has a particular orientalion, it maintains this orientation despite the S-spin pulses. Therefore, whether the I - S coupling enhances or diminishes the 5-spin precession rate does not matter. The dephasing produced thereby between the first and second pulse is refocused between the second pulse and the echo. Now consider the electron case. Suppose the electron (5) is initially pointing up. Let the nucleus (I) be pointing up along the effective field. Thm is, the nucleus is in a stale u+ of (11.90) and the total system is in the state 1+ +). The 1r/2 pulse now tilts the electron. As a result, the nucleus suddenly finds the effective ficld acting on it tilted. The nucleus thus begins to precess about the new field direction. Consequently the field it produces on the electron changes in time. Clearly the magnetic field acting on the electron during the interval 0 to T will, in general, differ from Ihal between T and 2T. Therefore. the echo will not refocus the spin-spin coupling completcly. If, however, T corresponds exactly to a 2'/T precession of the nucleus, then the spin-spin coupling will have the s:lme time average during the two intervals, hence the echo will be perfec!. Thus, we expect Ihe echo envelope to oscillate with T such that maxima in the echo envelope correspond to r's that are integral multiples of the nuclear precession period. We will see in our fonnal treatment of the problem that the description we have given is exact. In that connection, we can note that 10 think about lhe quantum treatment of the effect of a '/T/2 pulse on the electron system, the general concepts developed in Sect. 7.24 are helpful. In particular, they help us deal with the problem that there is not just one nuclear precession frequency. Rather, there are twO, depending on whether the electron is in its spin-up or spin-down state. We begin by expressing the reduced Hamiltonian (11.83) ill different notation. First we define the frequencies a, b, We, and W n :
rl2a :: Ax sin:! () + A z sin:! () h2b :: (A'l; - Az)sin () cos () W
n
::inHO
(11.92)
hWe :: 2(JHo Then 'Hu'<J ::
h[we5 z' - Will..., + (a1z1 + b1'1:' )25z/]
(11.93)
We now !Tansfonn this Hamiltonian to the electron spin rotating frame at frequency W
De:: We
-
(11.94)
W
using the operator exp(iw5z d) 10 !Tansfonn the wave function
t/J1::e-iwlS.,t/J
with
(11.95) 525
524
1i~ed '" fl[neS~, -wnl~, + (al~, + bl:t:,)2Sz']
(11.96)
To save in writing symbols, we now drop the primes from Xl and z' in (11.96), understanding that from now on the new unprimed axes are really the old primed axes. To solve for the eigenstates we note first that Sz commutes with 1i~ed' Therefore, the eigenstates, 1/J1 , of 1i:.w can be labeled by ms, the eigenvalue of Sz. Thus
1fred ¢'(ms) "'1i(OemS -wnl: + (al: + bIz)2ms]¢'(ms) = ~il.ms + 'li/(msl] .p'(ms)
For ms ::::
(11.98)
-! (11.100)
We now define two vectors w+ and w_ by
w+",j(-b)+k(wn-a)
(11.101.)
w_=ib+k(wn+a)
They enable us to write (1I.IOIb)
The vectors w+ and w_ are shown in Fig. 11.9. Their magnitudes are given by
z+
which give for the components of I along w+ and w_
,
1z _
'"
k_ . I
(11.105)
These enable us to write
::::-flW+·1=-/lW+1:+
1i,( -
i):::: -llw_· I'" -llw_lz _
I
Wn -
x
I
1z+u+ '" il.l+
1:+1.1_ '" -41.1-
1:_v+ :::: !v+
1,_1}_ '" -iv-
(11.107)
we have that the u's and v's are the eigenfunctions of 1t,(ms) as follows:
'liIWU+ 'li/
'" -/1(W+!2)I1+
'" 1I(w+!2)I.1_
'iiI ( - 1)"+ '" -li(w_!2)u+ 'iiI ( - 1)"_ '" -li(w_!2)u_
(11.108)
1++) =:(tu+
1+
1-
l- -) " pv_
+)
"pu+
-) =: (tIl_ (11.109)
that they are eigenfunctions of 1t~ed with eigenvalues
I
I
L_.J.:__
I (J
W+
h(ilo - w+)/2 , h(-ilo -w_)/2
I
1__ -"---_8+ -L-
_ x
FIg. 11.9. The two angular momentum vectors '"'+ and '"'_ in the :t: - : plane, defining the angles 8+ and 8_ made by '"'+ and '"'_ respectively with respect to the ~-axis. Note the sign conventioll (indicated by the arrows) for positive 8+ and 8_. The axes:+ and ,_ lie along '"'+ and '"'_ respectively 526
(11.104)
Then, returning to (11.97) to express 1i~ed' we find that when 1t~ed operates on any of the four functions
z
z
-b-
We then define the unit vectors k+ and k_ as
Then if we define the eigenfunctions 1.1+. u_, v+. v_ by the equations
'iiI ( -!) = -h[(wo + a)I. + bI.]
+a 6_
(lU03)
wn+a
(11.106) (11.99)
'li/m = -"(w" - a)I. - bI.]
Wn
b tao 6_ '" - -
Wn-a
1i,(!)
+! we have
and for ms '"
b tao 8+ '" - -
1z+ :::: k+ . I
where 1i/(ms) is an operator in the components of I, given by
1i/(ms) '" -1i(Wn1: - (a1: + b1:t:)2ms]
The directions with respect to the z-axis are shown in Fig. 11.9. The angles 6+ and ~_ shown in the figure obey the equations
(11.97a) (l1.97b)
•
(11.102)
Ii(nc +w+)12 , h( - ilo +w_)/2
(J 1.110)
respectively. The four functions of (11.109) therefore solve the reduced Hamil~ tonian. Utilizing the definitions of 6+ and 6_ given by (11.102) and shown in Fig. 11.9. we can calculate the angle 60 between z+ and z_. shown in Fig. 11.10. (11.111) 527
•
[ z
J.'ig.1I.10. The angle 80 between the angular momentum vectors \<>+ and ",_. With 0+ and (L both positive in the sense of thO) arrows, (}o = 8+ + 9_
where for simplicity we do not write out the contents of the inverse brackets and where the subscripts I and 2 on the brackets serve to distinguish the contents of the brackets. E.xplicit evaluation of X S1(1T)T(t - r)XS(rr) gives us X
s
1
(1t)T(T)X 5(11")
= exp{i[ .RcSz + (alz + bI>;:)2Sz + Wnl:] (t - r)}
(11.117)
We now note that we can commute the DeS: factors, to obtain
(S+(I) = ""--
~
Te{ S+ [Xs(.)"p{ i[D,(1 -
r) - D,r] S,}
x exp{ i [(aI, + bI.)2S, - woI,] (I - r) 1
x
xexp{ -i[(aI,+bI.)2S.-woI,]rl],SY[ Now, owing to the spread in De due
Referring to (11.89) and Fig. 11.8, we have that
¢ = -80
(11.112)
so thaI
10
(11.113)
We are now ready to consider the electron echo signaL We begin by assuming that Ho has some inhomogeneity. This may either resuil from actual inhomogeneity in the magnet or from some other source such as hyperfine cou-
pling to nuclei other than the spin I. This inhomogeneity would cause the free induction signal of the electron spin 10 decay, and thus causes us to go to a
magnelic field inhomogeneity, the
tenn exp{i[fle(t ~ ,) - fleT1S z} makes the signal {5+(0) be small at values of t. However, we note that when t '" 2t the factor
XsC1r)
(11.114)
Then defining the lime development operation T(t), in the electron spin rotating frame we have T(l)
=ex =
p(
maSI
Therefore, at t = 2. the signal is independent of fie, and an echo forms. We now confine our attention to this time. We next factor out X S (7r) from the left-hand side of [ h and from the right-hand side of [ J31, and utilize the fact that Tr{ABC} = Tr{ CAB} 10 produce X
1
(IU20a)
S (1T)S+ Xs(rr)
which we rewrite as
spin echo to eliminate the effects of inhomogeneity. We therefore analyze for the signal at time t = 2r for a pulse sequence T ...
(11.118)
(11.119)
v+ = u+ cos(80 /2) - u_ sin(8o/2) v_ = u+ sin(8o/2) + u_ c05(80/2)
XS(1rI2) ...
13'}.
(11.120b) Moreover, we get a more compact notation if we define the unitary operators R(Sz) and R(-Sz) by
R(S,) " exp{ -i[(aI, + bI.)2S, -woI,]r}
(11.121)
R(-S,) " exp{i[(aI. + bI.)2S, +w"I,)r}
(11.122)
,ad
-*1{~edt)
«p{ - i [D,S, + (aI, + bI.)2S, - woI,] I)
(11.115)
where R(-Sz} is obtained by replacing 5 z of R(S:) by -5:. Then we get
We take the density matrix e(O-) prior to the first pulse as ,q(O-) = Sz. Then the signal at time t (we take t to be after the 11" pulse), (S+(t», is given by
(S+(I)) = Te{S+ e(l)) = Te{ S+ [T(I -
r)X s (·)T(r)Xs(·/2)] , S,[
= Te{ S+[Xs(.)Xs'(.)T(1
We now evaluate the (face using a complete set of basis function. For the complete set we pick initially the two electron States a and fJ with quantum number
1,' 1
- r)Xs(.)T(r)],S,[
I,'} (11.116)
528
ms and the two nuclear states u+ and u_. which we denote by the quantum numThus, using the fact that (msllIS_lmsp./) = 0 ber p.. which can be +! or unless p. = l, and so on,
-!.
529
L:
(S+(2T) =
(msIIIS_lmSII)(msPIR( -S:)R(S:)lmSIt')
Then, for any function j expressible in powers of the components of I,
"'S,I'
...5·.. ',1''' X
R1/2fR_I/2=exp(iw+Tlz)jexp(-iw+rlz)=j
(msJ/ lS'y Im~Jt') (lnsp'IR-1 (Sz)R- 1(-S:)lm~plI) (11.123)
In order to have the sums be diagonal sums, we get that m~ = ms and p" = JJ. Now the matrix element of S_ requires that
ms=mS-I.
(11.124)
ms
-t '
=
+!
(11.125)
,. ,
and therefore (mspIS-lm's II) = 6", ",' -I' Then we get for Ihe matrix elemen! of Sy (ms/IS,lms/) = WS,I-!) = ;i WS+I-!) = ;i
so that
(S+(2T»
=
(11.126)
;i L (!,'IR(-S.)R(S.)IV) x (- !/IR-1(S.)n-'(-S.)I-
!")
(11.127)
Now, lcuing (he operator S: of R(Sz) and R(-S:) operate, we replace S: by ms (and -Sz by -ms) in the matrix elements. Then, defining the operators
R- 1/2 =exp{i(alz+bl.l"+wnI:)T} , = exp{i( - aI: - bI", +wnlz)T}
we get
(S+(2T»
=
,
~iTrJ{R_I/2RI/2R=:/2RI>~}
530
.
(S+(2T»
I I = 2iT
.
(11.134.)
(S+(O+»
s+
= T<{ S,) = iT< { S~) = i(2[ + I)Trs{ S~} . S(S+I) i =1(2[+1) 3 (2S+I)=2(2[+I).
(l1.I34b)
Therefore, if W+ T = n2:lf we get an echo corresponding to the full magnetization. A similar argument holds for the condition (11.135)
if we focus on the terms R_I/2RI/2R::/2_ We can get funher physical insight by writing
R_l/2RI/2R=~/2Ri/~ = exp(iw_ TI z_ )R 1/ 2 exp( - iw_ TI z_ )n-;/~
T
= exp(iw_ I z _) exp [ - iw_
T( RI/2Iz_Ri/~)] (11.136)
(11.137)
otherwise
,
(11.129)
R_ 1/ 2 =exp(iw_Tlz _)
(11.130)
Therefore, R 1/ 2 generates rotations at angular velocity w+ about the z+-axis, and R_ 1/ 2 gener.ates rotations about the z-axis at angular velocity W_, Suppose now that T corresponds to an integral number of 211" rotations about the z+-axis: w+T=n211"
(11.133)
Thus
R 1/ 2 exp( - iw_ rI z _) Ri/~ = exp( - iw_ T I z _)
where TrJ means a trace over the nuclear spin eigenstates only. We can express the operators R 1/ 2 and R_ 1/ 2 in a ronn that gives them a simple physical meaning now by llIilizing 01.101 and 105) ,
=1
Now RI/2Iz_R~/~ provides a rotation of the veclor k_Iz _ about the vector w+. When this rotation is an imegral number of cycles
;i L ("IR_1/2RI/2I''')(I''IR=:/2R~AI")
R t / 2 =exp(iw+Tlz+)
R-l/2Rl/2R=:/2Ri/~ =R_ 1/ 2R=:/2
(11.128)
/,,/,'
=
since. every component of I is rotated onto itself for this value of W+ T. Therefore
W_T = n211:
Il,Il'
R 1/ 2
(11.132)
But this is, apan from the usual sign change in a 1f!2, 11" echo, just the signal we would have immedialely after applying an XS(1I:!2) pulse:
Therefore, we get contributions only from
ms =
,
(11.131)
Rl/2R=;/2Ri/~"4 Ri/~
(11.138)
and the producl
R_1/2R1j2R=:/2Ri/2"41
(11.139)
Clearly these equations express mathematically our initial physical description of why lhe spin echo envelope is modulated as T is varied, returning to its full amplitude when either w+T=n211"
or w_T=n211'
,
n=I,2,3 ....
(11.140)
We now wish to evaluate the lTace explicitly. Denoting the eigenvalues of 1z + by p and of 1z _ by q so that 531
(11.141)
we get
TTl {R_l/2RI/2n-~
I::
=
," ,,11:,,11' =
L
(11.142)
e+ iUW _
E (T
TrdR_l/2Rl/2R=~/2Ri/2} = 2
:/2Rl>~}
(q IR_ I / 2 Iqll ) (0.1 IR 1/ 2 1qll) (u"in: :/210'1/1) (alII IR~/~ 10")
",17" =
T
(aIR 1/ 2 Iqll ) e-iUItW_T (qIlIR1/2 Iu)
eiW_T(
(0"IR 1/ 2 Iqll )
,U"
(a IR1/ Iu) Il
(11.143)
2
"II ,I''''
and relabeling of I~+, we get
{R-l/2RI/2R=:/2R0~}
=2
0'"
as (/ and using the faci that the states III) are eigenstates
11.5 V, Center
ei~.LT(O'-"") (ulp)eiw+TI-' (11.10-')(/7'1//) e-iW+T/1' (/l'la)
",u '
1'",1
(11.144)
Since 17, ql, J~, and fJ.' can each take on two values (+! or lenns. The factors (ulp.) are given by expressions such as, if
17=! (ulr) =
.
Thus, we get the full signal, as we showed with (II 133).
TTl {R_l/2R'/2R=:/2Rl/~}
-!) there are
16
fl=-!
Jv~u+d7"
(11.145)
where the integral is easily evaluated using (11.109):
(0"::: ~I/~ = -!)
= sin(Bo/2)
(11.146)
and so on. It is then straightforward to show that
Tr I {R_I/2Rl/2R=:/2Ri/2} 2 2 = 2 [ cos 4(00/2) + Sin 4(00/2)] + 4 sin (00/2) cos (00/2) x [COS(w+ ,) + cos(w_ ,) - cos(w+ ,)cos(w_ ,)] 532
cos(w_,) = 1
2 2 = 2 [ cos «()o2) + Sin «()o2)t
x (fJ. "IR-II 1/2 I' "')( II- "'I (1 )
L
Tr I
or
4 = 2 [ cos «()0/2) + sin 4 (00/2) + 2 sin2«()0f2) COS 2«()012)]
,,',," 1',1"
,
which 'agrees with (11.134). That is, it is independent of ,. Thus, the oscillation comes about because of the change in quantization direction of the nucleus when the electron spin flips. Note also that if either cos(w+,) = I
=
=
If 00 = 0, so that {here is no change in the quantization direction, (1 t.147) reduces 10
(11.147)
A particularly interesting example of the application of the ideas of the previous sections is the discovery and identification of the so-called Vk center by Kiinzig and CaStner (11.5]. The detailcd analysis of the spectra will enable us to discuss the g-shifl more thoroughly, including the effects of having several electrons and more than one force center. Kdnzig and Castner's first work in electron spin resonance was followed by a set of beautiful experiments by De/becq [11.6], who combined optical techniques with electron resonance to detennine (I) the optical absorptions associated with the center and (2) the energies of the excited stlltes. A full account of all the work on Vk centers would take us too far afield. We shall remark on the method of identification of the center, and on certain features associated with the g-shift, which are not found in "one-atom" one-electron centers. It is helpful to begin by, so to speak, giving the answer. The Vk center is formed in alkali halides by X-raying crystals al a temperature near that of liquid nitrogen. In this process, electtons are ejected from the negative halogen ion, changing it from a closed-shell configuration to one with one electron missing from the p-shelJ. The ejected electron may have a varicty offatcs. We shall simply assume that not all recombine wilh neutral halogens. As an example, consider CI; the neuttal chlorine atom is unstable and pulls togcther with a neighboring Cl- to form what may be convenicntly called a Cli" molecule. The Cl-CI axis turns out to lie in 110 or equivalent crystal directions, as illustrated in Fig. lUI. 533
+
+
+Y:J
+
+
+
1'lg.IUI. A eli" molecule, or V. center. in KC!. The
«nl~
be thought of
pair of Cl-
ILS
a hole, denoted by h, trapped on
Il
may
ions
i.
((/+
+
recognized to arise from the coupling of the unpaired electron spin with magnetic moments of a pair of chlorine nuclei. Let us explain. There are two isotopes of chlorine: C135, which is 75 percent a~undant; and C~37. 25 percent abundant. Both are of spin bUI have slighLly different magnelJc moments ("(37/7'35 = 0.83). As we shall show. the seven-line pa~lem arises from a pair of Cl35 ·s. Let us assume that Ho lies along a principal axiS of the eli; call it Ihe z-axis. By recalling our discussion of the preceding section and generalizing it slightly, we see Ihat the electron resonance condition will be
The electronic sttucture turns out to be very similar 10 (hat of the p-electron with quenched orbital angular momentum, discussed in Section 11.2, the electron with the unpaired spin being in an orbit whose axis is parallel 10 the bond direction of the eli" molecule. Coupling to excited stales gives a g-shift thai varies as the magnetic field orienlalion is varied with respect 10 the crystal axes, and the coupling of the unpaired spin with the nuclear moments of the two chlorines gives rise to a hyperfine coupling. As we have remarked. this center was first discovered by elccuon spin resonance. A pattern observed for a case where the stalic magnetic field is along the 100 crystal direction is shown in Fig. 11.12. At first sight the spectrum seems too hopelessly complicated to unravel, but fortunately a convenient starting point is the set of seven prominent lines. They are nearly equally spaced and are of intensities 1: 2: 3: 4: 3: 2: I. These were
1
w = h[9uPHO + A,(m) + m2)]
where m, and m2 are the m-values of the two CI nuclei. Therefore the frequency depends on m, and m2 only through their sum ml + m2. The largest m, + m2 is + ~ = 3. The next largest is + = 2. This value of ml + 1»2 is also found if mt = ~. m2 = Since we assume the nuclei to be distributed at random among their m-Slates, the line at m) + m2 = 2 will be twice as strong as thai for m) + m2 = 3. We indicate the possible combinations of mt and m2 in Table 11.2,
i
! i
!.
The recognition Ihat the seven main lines could arise from a scheme such as we have described was the first clue as to the nature of Ihe resonance. If one accepts this clue, Ihat Ihe electron spends equal time on two CI atoms, one mUSI next consider what happens for CJ; molecules in which the nuclei are bolh 35 and the other is a C1 37 . The probability of finding a Cl37,S or one is a 3 Therefore the probability single atom of CI 5 is ~; that of finding a CI37 is of finding pairs is as follows:
a
~ 0 10
a 35 _ CJ35:
20
CJ35 _ CJ37:
~30
CJ37 _ C135:
~"
Cl 37 _ C137:
"
I
II.
1
II'
" 70
80 90
100
,
. ,
,
• ... ,
H
,I
1
'I
I
3.3 ,
,
U,
,
.. ,
*
[9uPHO + (A;5 m35 + A;7 m37 )J]
(11.149)
in an obvious notation. The hyperfine coupling, as we have seen, is proportional to the nuclear "('s, so that kilocauu ,
Fig. 11.12. VI: rellOnance in KCI for ~he static field parallel to a 100 crystallographic axis. (This figure kindly supplied by Klinzig and Casfnl!r)
534
t.
The generalization of (11.84) for unlike nuclei may be wriuen as
w=
,
(11.148)
A
3S
7'35
A37 = "Y37
(11.150)
The result of having nonequivalent nuclei is that configurations such as (3 I) t
3
.
~'~
and (~.~) no longer gIVe the same frequency. If we call the (i.~) intensity 53S
Table 11.2. Combinations (ffil,ffi7) to make the same total valuc m\ corresponding frequcncy and statistical weight
(~,~) (~,~)(~,~) (:I.2' ~1)(.1 1)(_.1 :I.) 2 7'2 2'7 (;l _i!)(1 _1)(_1 1)(_i! ;l) 7'
2
2'
7
2'2
2'2
(-~, ~)( -~, -t)( 1, -~) (-t,~t)(-t,-~) (-~,-~)
3A 2A
1
lA
3
OA
4
-IA -2A -3A
3
+ ffi7,
and the
2
2
(11.151)
"unity", the line that for like nuclei would have intensity 2 splits into two lines of unit intensity when the nuclei differ; that of intensity 3 splits into three lines of unit intensity, and soon. The positions are all predictable, using the measured A~5 and (11.91). The intensity of the 35-37 lines is that of the outennost line 37 of the 35-35 spectrum. There is in addition a set of seven lines from the Cl 35 pairs. also having a predicted position, with an intensity ~ that of the Cl pairs. All these lines are found at the proper positions and with the proper intensity. In this way many of the lines are accounted for. An additional factor in determining the spectrum is the fact that the g-shift displaces the position of the center of the hyperfine panems. The principal axes of the g-tensor with respect to the molecule are shown in Fig. lI.l3. As we see in Fig. 11.13. if the magnetic field were perpendicular to the plane of the paper (pointing in a (001) direction), it would be parallel 10 the y-axis of the center. IC it were along the 100 or 010 directions, it would make an angle of 45° with respect to the z-axis of the molecule in Fig. 11.13. For any given orientation of the magnetic field with respect to the crystal axes, there are in general several classes of Vk centers in tenns of the angles made by Ho with the principal axes of the center. If H 0 is parallel to a 111 direction. there are two classes of molecules. If HO is parallel to the 100 direction, of the centers have their bond axes perpendicular to Ho and ~ have a bond axis making a 45° angle with Ho·
t
l
, +
+ + 536
+
We can see that in general there will be several hyperfine patterns whose centers of gravity are displaced because of the anisotropy of the g-factor. Moreover, the hyperfine splitting is itself strongly anisotropic. A~ being much bigger than either A:t" or. A y (z being the bond axis). The anisotropy is interpreted as indicating thal on each atom the individual bond function is a linear combination of an s-function g(r) and a p-function zj(r).3 The hyperfine coupling becomes, then, using (11.65) and (11.66),
Fig.II.B. Principal axes :1:, y, ~ of the g-tensor for the V k center. The y-axis points out of the paper
where 0 2 is the fraction of p-function, (l/J.J) the average of (l/r 3) for the state zj(r), and where a factor of multiplies the expression of (11.67), since the wave function spreads over two atoms. (We have neglected renonnalization due to overlap of atomic functions.) There is a near cancellation of the two terms multiplying (II + 12)·5, leading 10 a strong anisotropy. The fact that the electron ranges over more than one atom presents a new problem in calculating the g-shift. In the example we slUdied, we represented the spin-orbit coupling by a form >'L· 5 appropriate to a free atom. The origin about which the angular momentum was measured was, of course, the nuclear charge, since it is motion of the spin with respect to this charge that produces the spin-orbit coupling. When there is more than one nucleus, it is not apparent which nucleus to choose as the origin. The dilemma is resolved by using the more basic fonn of the spin-orbit coupling:
4
eli.
'HSO=-2 22S.(Exp) (11.152) me In this expression E is the electric field through which the electron is moving, and p is the momentum operator of the electron, (M)'V. Since for an isolated atom E is directed along Ihe radius vector from the origin, Ex p ex: r x p, the angular momentum. The usual >.L . 5 expression therefore follows from (11.152). Since the electric field E is largest near a nucleus, the principal contribution to 11.50 comes when the electron is near a nucleus. In addition to the presence of two force centers, we have the complication that we must deal with more than one electron. In fact the Vk center lacks only one electron to fill the valence shells of its two chlorine atoms. In order 10 proceed funher, it is helpful to describe the electronic states. We shall describe
3 The electronic wavc functions are discusse<1 in the subsequent text. We may remark that the bond function refcrroo to here is the function ~\ + ~2 mentioned in the latcr discussion.
537
...,, , /'
.....
"-~
.- '
,,~ ~/
+
Fig. 11.14. Functions (a) %1 %2 and (b) %1 -':2 shown 5Chematiclllly. The function of (a) is higher in energy, !lincc I) it hll.!l more nodCllllnd 2) the node lies ftt a region of aUractive potential for the e1eclron
is from the zi + Z2 to %1 - .1:2, since the electronic dipole matrix element here is the largest (in fact, corresponding to a dipole moment arm equal to the length of the molecule). We see, therefore, Ihat we muSt generalize our previous discussion to account for two new features: the lack of a single force center, and the fact that more than one eleclrOn is involved. In order to illustrate the first point (more than one force center) without the complications of the second (more than one electron) let us consider an example in which we have only one electron occupying the VA: center orbitals. Then the ground orbital state function !Po is, neglecting overlap, 1
(11.153)
!Po "" /2(ZI - .1:2) them in a molecular orbital scheme in which the molecular orbitals are made up as linear combinations of the free-atom p-states. We shall denote by :1:1 an atomic p-function zf(r) centered on 3tom 1 of the center. The atomic functions are thus %1> YI> zl. %2. Y2, and %2. the z-axis lying along the bond. The functions %1 + %2 and %1 - %2 are shown schematically in Fig. 11.14. A study of the figure shows that %1 - %2 corresponds 10 a lower energy than does %1 + %2. since it has fewer nodes and lends to concenlrate the electronic density between lhe two atoms where it can share their altractive potential. The Slates are in fact referred to as bonding (ZI - Z2) and antibonding (ZI +Z2). In a similar manner it turns out that XI + X2 and YI + Y2 are bonding, and XI - X2 and YI - Y2 are amibonding. (The z-states are the so-called q-states and the X or y states are the lI'·states.) The energy levels of these states are shown schematically in Fig. 11.15. Actually the states Xl + X2 and VI + Y2, which are degenerate in the free molecules, are not degenerate in a crystal, but we neglect that splitting. Since there are 6 orbital functions, there is room for 12 p-electrons. The VA: center, which has only II, therefore has a hole in the ZI + Z2 state. That is, there is an unpaired electron in that state. We have introduced, in Fig. 11.15, a labeling u or g (ungerade or gerade) that describes the parity of the orbital state. One may expect to observe an optical absorption due to VA: centers. Since electrical dipole transitions are allowed only for transitions u to g or g to u, the optical absorption will arise from transitions of an electron in either the states ZI - Z2, Xl - x2, or YI - Y2 to the empty %1 + %2. The strongest optical transition
538
"1
t
" - '2 (u,)
,
(r,,)
(11.154)
m ,
In the absence of an applied magnetic field this expression is correct but in the presence of a magnetic field described by a vector potential A we must modify it to oblain a gauge invariant result: 1ioo ""
2~~c2S'
[E x (p+ ~A)]
(11.155)
where -e is the charge of the electron. Equation (11.155) follows directly from the Dirac equation but it is also intuitively obvious because. as we have discussed earlier, we always replace p by p - (qlc)A in the presence of a magnetic field where, in our case q -e. The orbital Zeeman interaction 1ioz is
=
2mc
2mc
(11.156)
We can consider that 1iOZ and 'HSO together constitute a penurbation 1iperL :
(11'",) "y, --• "J Y2
"y, + "2
eli.
'liSO=-2 "S·(Exp)
" A2 1ioz "" -2( p ' A + A· p) + __ 2
'. + .1:2 (u,,) ,,
which has quenched orbital angular momentum. In our earlier discussion of the g·shifl for a problem with only one force center we saw that the g-shift arose from the interplay between the spin-orbit coupling and the slight unquenching of orbital angular momentum produced by the orbital Zeeman energy. When more than one force center is present there is no unique point about which to measure angular momentum and it is natural to return to the more basic fonn of the spin-orbit coupling:
1iperL ""
Fig. IUS. Molecular orbiUlIs ronned rrom ,..sta~ in .. halogen molecule ion. The allowed optica.ltransitions into the unfilled <1'" orbit &reshown by dashed line!l. The ll'..hil'l "transitions" an shown by a solid line
1iso + 1ioz
(11.157)
We are concerned with calculating maaix elements designated by an orbital quantum number n and a spin quantum number o. In panicular the effect of1ipert is, according to Appendix D, equivalent to our having an additional interaction 1in_ whose matrix elements diagonal in the ground orbital state 10) are 539
where 1
(11.158) where the prime on the summation means omitting n = 0 and where we have neglected the spin contributions 10 the energy denominators. 11te g·shift arises from keeping just those terms of {I 1.158) that are linear in the vector potential and the electron spin. Thus we get (Oo"l'H.1g[OoI) as
(Ocr l1i<191 0a/ ) =
2~~c2 (oaI5.E x~AloO"I) e2h
~I (OaIS·E x plnull)(nq"lp·A +-3 -~ L.J n17" E0- En 4mc-
(00'11" A + A . plnu'I)(nu"IS. E x 1'100') + Eo-Ell
+ A'pjOql) . (11.159)
e2h
2m'c3 (aISla') (OlE x A(R)IO) 0
e2h = 4m'c3 (aISla'HOIE x [lio x (r - R)JIO)
(11.165)
(11.166) Then taking the z-axis to a lie along the direction of the stalic field, we find
,'h
2~2:3(qISI0'1). {(DIE x AID)
2m'c' (aISla') (OlE x AIO) 0
+ _1_ L,[(OIE x pln)(nlp A + A plO) +c.c.]} 0
2m n
0
Eo-En (11.161)
This expression is the basis for a proper treatment of the problem of several force centers. However, in order 10 proceed wilh thai problem, we should first understand some aspects of the single force center problem that we have not discussed. In particular, what is the best choice of the gauge for the vector potential, and what happens when we change gauge? Suppose the atom in question is located at the origin. Then the wave functions In) are in general either classified by angular momentum about the origin or are perhaps linear combinations of such atomic orbitals. If we took as the vector potential A(R) defined as
A(R) " 11Io x (r - R)
(11.162)
where R is an arbitrary constant vccwr, then, using the fact that div A(R) == 0, we could write the matrix elements (nip' A + A· pIO) as (nlp·A+A,pIO) ==
jU~Hox(r-R).puodT
==nHo' ju;,L(R)uodT 540
is the dimensionless angular momentum operator about the arbitrary point R. Integrals such as the lower one on the right of (11.163) are readily evaluated by the methods of (11.17) and (11.18) provided R is chosen as zero so that the angular momentum is measured about the natural origin of the atomic orbitals thac make up the functions In). We shall call this the "natural" gauge. An even more imponant point is seen by examining the first-order tenn of (11.161),
(11.160)
Therefore we gel
(OO'I1-l Llg IOa') =
(11.164)
1
By using the fact thai the electric field E is large only near the nucleus where it is in fact to a good approximation radial we have
Since p' A + A . p does not depend on spin (nul/lp, A + A . pIOu') = (nip, A + A . pIO)50'.0""
L(R) " 7(r - R) x 'V
(11.163)
= Ho
4;::~ (0'1510")' { (01 E;r) [k(x 2 + y2) -
- (OIE;'O)[k(XX+YY)_iX,_jY,llo)}
ixz - jyz
llo) (11.167)
where X and Y are two of the components of R. If the wave function 10) has a definite parity the second term on the right vanishes. If 10) does I10t have a definite parity (as for example if it were an .s-p hybrid) lhe second tenn does not vanish. Since this tenn depends on the choice of R it can in the latter case be made to take on any value. In order thac the g-shift be independent of the gauge, there must be a compensating change in the tenns of (11.159) that have the energy denominators. Such is in fact the case. If we take the "natural" gauge for which R == 0, the ordcr of magnitude of the right-hand side of (11.167) is approximlltely fiHoro/all' where 1'0 is the classical electron radius e2/mc 2 (~1O- 13 em) and all is the Bohr radius (0.5x 1O-scm). The matrix element is therefore S!' 1O-sfiHo and is in gencral negligible. It is for this reason that one is justified in omitting the first-order lenn, as is ordinarily done. We have seen thai the "natural" gauge makes for simple evaluation of the matrix elements such as those in (11.163). When there is more than one force center no single gauge appears natural, and we should like, in fact, to be able to 541
use a mixture of gauges: one gauge when in the vicinilY of one nucleus, the other in the vicinily of a second nucleus. Such a trick is aClually possible, provided we can neglecl cenain overlap integrals. Lei us Slate a theorem; lhen we shall outline ils proof and then show how Ihe lheorem enables us 10 use such a lechnique of several "natural" gauges to treat the problem of muiliple force centers. Let us therefore consider a system with two atoms. The ground stale 10) will be a linear combination
(11.168)
100=uo+vo
where uo is a linear combination of atomic orbitals on Ihe first atom and Vo is a linear combination of alomic orbitals on the second alom. The excited stales In) are also linear combinalions:
(OlE x A 10)
2~Z~ (uI5Iu'). {(UoIE x A'luo) + (VOlE x Aillvo) + _1_ E' (OlE x pln)[(unlp· A' + AI . pluo)
(Oul1ll1g l{)q/) = +
Eo En (un Iv' A" + All • vivo)] + c.c. + Eo _ Ell + (volE x V,plvo) 2m n
+ _1_ L;' (OlE 2m n
+ _1_ '£' (OlE x pln)[(uII1A' . p + p' A'luo) + (vnIA" . P + p' AI/lvo») + c.c.} 2m n Eo-En (11.170)
Ju~(A'
.p+p.A/)uodT
Ho
x (r -
Rt>
All = A2 = tHo x (r - Rz)
Eo
c.c.}
E"
eZh { (0.I H .o,(1)lvoo') " - 2 2 3('1 5 1")' (OlE x "71I vo)
mc
c.c.}
+ _1_ L;,(OIE x vln)(nlv' "71 + "71· vivo) + 2m" Eo-En =0 (11.176)
(11.171)
The integral 1, defined as
1==
Jtb~(p·V,p+V¢·V)fJOdT
(11.1n)
can be transfonned, making use of Ihe fact that the wave funClions are real and utilizing panial integrations, to be
(11.172)
To prove the Iheorem of (11.170). we stan wilh (11.161). We express the matrix elements involving A in tenns of the u's and v's. and neglect overlap tenns. For example, 542
vln)(v"lv' "71 + "71' vivo) +
To derive our theorem, (11.170). we must show Ihal the lenns involving ¢ add to zero. By making use of the faci Ihat we are neglecling overlap [enns, our proof is equivalenl to showing thai the quantity (Oul1ll1g(¢)lvou/). defined below. vanishes:
(they
The beauty of (11.170) is that il allows us to choose the veclor polential A' used to evaluate the integrals with Ihe u's independently of the veclor potential All used for integrals involving v's. C'Ne shall discuss handling the matrix elemems (OlE x pIn) shonly.) In panicular, we shall see, if the IWO nuclei are at R I and R z, respeclively. we can evaluate the matrix elements readily by choosing
A' = Al == tHo
X
(11.175)
2~zhc3 (qJ5Iq')· {(UoIE x A'luo) + (volE x A"I1l1J)
==
(11.174)
defining Ihe funclion ,p. (That (11.174) is simply a gauge transfonnation follows, of course, from Ihe faci that it satisfies the requirement V x A' = V X A"). We then substitute A' for A in integrals involving u's. and All + V,p for A in integrals involving v's. By collecting tenns. we gel
(OaIH.ogIOo-')
(uIlIAI.p+p.A/luo)
(11.173)
AI = A'I + V?
We shall neglect all contribulions 10 matrix elements involving a produci of a u and a \/. This approximation is often good, but can lead to errors in some cases. We have, then, as our theorem Ihat the combined effeci of the spin-orbit and orbital Zeeman coupling is to give a g-shift characterized by
where A' and A" are any vector potentials Ihal give the static field differ therefore, at most by a gauge transfonnalion), and where
.
Then we introduce two vector polenlials. AI and All, which differ by a gauge transfonnation:
(11.169)
=
=(volE x Alvo) + (volE x Alvo)
II
1= ih
J
is simple
10
(vo¢V 2 1/Jn -1/Jn,pV 2 vo)d,.
(11.178)
re-express Ihe firsl lenn on Ihe righi, since
2 2m V tbn = /;2(V - En).pn
(11.179)
where V is the potential aCling on the electron. We evaluate the second tenn by 543
nOling that, neglecting overlap,
J
Wnt/J\l2 vo dr ""
J
v,,4> V2v odr =
!
2 vn4>V ¢odT
(IU80)
By utilizing (11.179) in (11.180) and again neglecting overlap, we obtain finally 2m (11.181) (nip' \74> + \74>' vIvo) = -:-. - , (Eo - En) 1/Jn¢vodT
j
r,
, r.
Therefore. when the upper sign applies, the two tenns in the square brackets of (11.185) cancel, and (DIE x pIn) vanishes. On the other hand, for the lower sign, the tenns add, giving twice either one. Thus the state (Xl + x2)/.,fi does not contribute to the g-shift, but the state (XI - X2)/,fi does. Of course a similar argument shows that (Y1 + Y2)1.,fi also makes no contribution although (YI Y2)1V2 does. The states involved in the g-shift are shown in Fig. 11.16 by the solid arrow.
We can substitute this expression into (11.176) and collect tenns to obtain
,'I.
[
(Oo'l'HLlg(t/J)lvoql) = 2m 2 c3 (uISI0"') (DIE x 'V¢lvo)
- 2 L'(OIE
x ''In)(nl''lvo)]
(11.182)
n
The prime can be removed from the summa{ion, since the diagonal spinorbit mauix: elements vanish, giving
(Oul'H"g(,,)lvou') = =
e. 2 tt
--(uISlu'). [(DIE x"" 2m2 c3
2,.
.
2(E
-~(O"ISI17I). jE x V(¢va)dr 2
x ""lvo)J (11.183)
2m cwhere (E x \7) signifies thai E x \l is to operate on all functions to ils right, that is, on both'" and VQ. But the integral can be shown to vanish. utilizing the fact that V x E = 0 and transfonning the integral J \l x (E¢wfi)dr into a surface integral. We omit these details since they are quite standard. Our theorem is thus proved. We have not as yet said anything about the spin-orbit matrix elements to excited states. By utilizing the fact that the electric field E is large only near the nuclei, we are always able to neglect overlap when evaluating spin-orbit matrix clements. Thus (DIE x pin) = (uolE x plu,,) + (volE x plv,,)
(11.184)
To see the full import of (11.184) as well as to illustrate our theorem (11.170) concretely, we now tum to the evaluation of the specific problem of a molecular complex in which only one electron occupies the Vk center orbitals. The ground state is therefore given by (11.153), t/;o = (1/V2)(zl - Z2). We are concerned with excited states such as (l/V2)(Xl ± X2). We have. using (11.184),
(DIE x pin) = (
Z1-Z2
.j2 IE x pi
= WztlE
.j2
(kI2~~"S'EXPll) =A(kIL·Sll)
(11.185)
(11.186)
(11.187)
where k and I denote free-atom states associated with the particular A. For our example. XI, YJ, and Z1 are being taken as free-atom p-states. Therefore we can write
'I., , (0-1510-') . (z11E x plxd = A(uISlo-')· (zIILllxl)
2m ,
(11.188)
where IiLl is the angular momentum about the nucleus of the first a1om, and where A is the spin-orbit coupling constant appropriate to the (np) electron configuration of the outer electron. Evaluation of the matrix element (zJlLJlx n ) proceeds as in (11.17) and (11.18). We now tum to evaluation of the matrix elements (unIAI. p+p .A'luo) of (11.170). We have that Un = xl/..fi or Yd..fi. uo = zd..fi. By utilizing (11.163) and the fact that the u's are real, we have (unIAI·p+p·AJluo) = IiHo'
J
UnLluodr
= -2-' (:1:1ILdzl)
Since the two atoms are identical and since E is large only near a nucleus,
544
If we had a free atom, the spin-orbit matrix elements could be expressed in tenns of the free-atom spin-orbit coupling constant A according to the equation
hHo
X1±X2)
x plxt) Of (z,IE x plx,)J
I.jg.ll.I6. The solid line indicates the states joined to the ground state Zl - Z2 by the spin-orbit coupling
or
But. by the symmetry of the atoms, (unlA1 'p + P ·Alluo) = (V,dA2' P + p' A2lvo)
(11.190)
so that we have. neglecting the first-order tenns such as given by (11.170), 545
(o.I1i",IO.') = 2PA(.ISI.'). [(Z1l
L 1I<1)«II L 1I ZI)
E:"I_z~ -
+ (z1I L 1IYI)(Y1I L ll z E Z \-Z2 - EY \-1/2
E:l:1-2:2
il). Ho
@
(11.191)
This is equivalent 10 having
L
1{i1g =
2PSqoqq,Hq,
(11.192)
(a)
qq'=z,Y,z
By evaluating the matrix elements, we get that Oqq' = 0 if q
+q'. and
Fig.ll.I& (a) Current now produced by the exurnallldd flo in the molecular complex. The fact that no current cru;ses Lhe boundary bet""een the t",'O atoms results from the neglect of overlap. If overlap i. included, the pallern is shown in (b)
A
°;t:z = -=-_.::...~- EIII -1/1 - E 2\-22 allY '"
"
(11.193)
-=--c...,~-
azz "" 0
E:l: 1 -:l:2
-
(b)
E ZI -'2
.
It is interesting to comment a bit more on why the states <:1:1 + %2)/.,fi and (Yt + 'J2)/..!i do not come into the g-shift. We note not only thai the spinorbit malrix elements to those states vanish, bUI also that the orbit-Zeeman lenns cancel. Mixture of these excited stales corresponds to production of a current
flow in the ground stale, shown in Fig. 11.17.
coupling, let us assume that we can characterize the multi-electron states by a total spin quantum number 5 with eigenvalues M for some component. An extra quantum number n will also be needed to define the energy. We designate the ground state, then, as 105M) and excited states by In5'M'). We are concerned as before with the spin-orbit and the orbital Zeeman couplings 1tso and 1toz, respectively. The same expressions will apply as for the case of one electron, except that we must now label the coordinates by a symbol j, to specify which of the N electrons is involved. We have, then,
l:1i~
1iso =
(11.194)
j
where Fig. I1.l7. CUlTent now produced by mixing some of the function (ZI + :1:2)/../2 inlo lhe ground !llate
(Zl - 21)/../2
[E' x (p, + ~A-)] _e_(p .. A-+A ..
~S"
1i(;) so = 2m2c2) 1i(;) =
J
J
C
)
(11.195)
p.) OZ2mc))))
According to (11.155), the gauge-invariant spin-orbit interaction is between the spin and the gauge-invariant current density j(r). With a current flow such as that given by Fig. 11.17, flowing in opposite senses on the two atoms, the net spin-orbit coupling vanishes. That is the significance of the vanishing of the spinorbit matrix element. The vanishing of the orbit-Zeeman tenns represents the fact that the applied field would never induce a current flow oppositely directed on the two atOms. Rather, it gives a flow such as that shown in Fig. 11.18. We see that the technique of handling the problem of several force centers is to break the integrals into tenns that are large only near the individual force centers, thereby converting the problem to a sum of single force center problems. The second problem with which we must grapple in order to analyze the VA: center is how to compute the g-shift when we have a system with more than one electron. Since the spin and orbit are uncoupled in the absence of spin-orbit 546
and where we have neglected the tenn involving the square of the veclOr potential in the orbit-Zeeman coupling because we seek tenus linear in H o· For simplicity we divide the spin-orbit coupling into twO tenus, one involving the vector potential A; the other, not
e)
J'J 1t(SOA =2m2c2 -d, - 5 "J ( E·x-A· J c)
(11.196) 'H.(j)
- ----=!!"-S· ·(E·
500 - 2m2c2
J
)
x p.) J
Therefore 'H.soo is the spin-orbit operator in zero-applied field. Assuming that the orbital angular momentum is quenched, we have, therefore, that
(OSMI1iozIOSM') =
°.
(OSMI1isooIOSM')
.
(11.197) 547
Then the spin-orbit and orbital Zeeman coupling combine to give matrix elements equivalent to OUf adding a tenn 1(. 6 9 to the Hamiltonian, where
we get
(OSMI'Ii"gIOSM') =
(OSMI'Ii"gIOSM') = (OSMI'lisoAIOSM')
+
:L
(OSMI 1isoolnS' M")(nS' MlIl?iozlOSM')
nS'M"
Eo-En
j"inlosM) '''';11 105M')
Eo -
En
(11.199)
.
Since the total spin of the VI; center is ~. the
I
11) I....
0'2'2
Siale
105M) =
IOi-l) is then
p
It~. +~1,+(3) ... :. +:1,+(11)
(11.200)
That is. all the orbitals except u:. +:1,- are occupied by an electron. It is convenient to denote functions such as (Zl - Z2)/,fi by a symbol I and the spin quantum number by (I (since m is also used for the mass of an electron). In this notation the individual electron orbitals are denoted as 1/0). As discussed in Section 4.9, all matrix elements of (11_198) arise from one-electron operators, so that the states joined by the operators can at most differ in the occupation of one orbital. Thus we find that we can express (05MI1tsoAI05M') in terms of the one-electron operator
1i~gA:
(OSMI'lisoAIOSM') =
L: (l.I'Ii~gAI/.,)
(11.201)
10' ,/0"
where 10 goes over all values occupied in 105M) and where la l goes over all values occupied in 105MI ). We do not include matrix elements (lal1iSOAl/lal) where zt i- I, since these states imply a change in the occupation of molecular orbitals. That would imply that the ground state possessed orbital degeneracy, a circumstance we do not wish to consider. The second-order terms are handled in a similar manner. The sum over electrons can be converted to a sum over orbitals occupied in the ground state and sum over n to a sum over orbitals not occupied in the ground state. Therefore 548
(lal1t~~ 11'0')(/'a'l1igil/a')
L:'
+
~LrkLion
El - ~ll
B
E,- E" (11.202)
where by restriction A we mean:
110) is occupied in 105M) 110') is occupied in 105M')
Ita)
is occupied in neither 105M) nor 105M')
and by restriction B we mean:
1/0) is occupied in 105M) 110') is occupied in 105At) Ita') is occupied in neither 105M) nor 105M')
:: /iTf'''p)-I) PU':'_':l,+(I)u':'_':l,_(2) X
(/.1'Il!OZ I) 1/'.)(1'.1'Ii[I) II.') SOO
resuictiol1 A
(11.198)
where we have neglected the spin in the energy denominators, and where we are also keeping only the terms that give rise to a g-shift. In the case of the Vir center, we may take the wave functions 105M) 10 be a product of one-electron molecular orbital states, properly antisymmetrized. The calculation proceeds along lines similar to thai of the indiIttt nuclear coupling in Section 4.9. Lei us denote the state (Zt - %2)/../2 containing electron number 1 with spin-up (m = +!) as u,,-.z=J,+(1)
L:'
+
+ (OSMI1fozlnS' M")(nS' M/lI11'sooIOSM')
Equation (11.202) will hold for any system in which the wave function can be ta);:en as a product of individual spin functions. We wish, of course, to incorporate our earlier theorem, (11.170), to enable us to use "natural" gauges. This can be done readily by noting two things: The first is that, since 1lgi is independent of spin,
(l'I'lig~lt.) = (l1'lig~lr) = (/"I'lig~I'.,)
(11.203)
By utilizing this fact, we can see the second point. If we remove the condition that II' a) be unoccupied in 105M) or 105M') from restriction A, and that Wa') be unoccupied in 105M) or 105M') from restriction B. the extra terms we acquire will exactly cancel in pairs. We can therefore write
(OS M 1'Ii", lOS M') =
L: (l'I'Ii~gAII.')
10';/0"
+
L: /(1;10";1'
1 ) II.') + (/.11i(seo (I. 11i(oz I) 11'.)(/'.I".A I) III a l )(I' a' I1t(l) '~SQQ oz II.')
E, - E"
(11.204)
where now we require only that I/a) be occupied in 105M) and 1/0') be occupied in 105M'). 549
where the components Sq are. for example,
Consider now all the teoos of fixed lu and lu':
N
(I) , ",(lql1ig~II'q)(I'ql1i~\5,IIu') (lql1isOAllq) + L, E E, I'
I
Sz =
E,-E"
(11.210)
etc.
i=l
I
+ (lql1i~&ll'ql)(ra'l1{gill<"')
L Sz;
(11.205)
This is identical in (ann 10 (11.161). II can therefore be convened 10 the expression involving the mixed gauge. Let us therefore define
For the VA: center, however, rather than employ the Wigner-Eckart theorem. we shall simply evaluate (11.208). Symmetry tells us that the principal axes of the gtensor are the x-. y-. z-axes of the molecule, where the z-axis lies along the bond. Suppose, therefore, that the static field lies along the x-axis and that M is taken as the eigenvalue of Sz. Since this is a principal axis. the only nonvanishing matrix elements have M = M'. Of course we can verify this by explicit evaluation of (11.208). S is, of course, Let us compute (ll.208) for M M' The simplest way to discuss the matrix elements is, then, in terms of a diagram of the states. Since we are computing a diagonal term in (OSMI1t.dgIOSM'). the states labeled 110') and 110") of (11.208) must be identical (we must relurn the electron to the state from which it was vinually excited). Setting 0' = 0 and neglecting the terms involving (luJV(l)llu), we have
4.
and
(lu!U(1)lt u') = 2:2nc3 (uIStlu')· [(udA'l' PI + PI . A'duII ) +(v/IA'i,pt +Pl,A'{lv/,»
(11.206)
In tenns of these definitions we can rewrite the expression of (11.205) as (I)
(lqlV
,
Ilq) +
E
+
I'
Ir
(lql1i(I) II'q')(YqW(I)IIu') E,-EI'
I: (lqlV(1)llq') IUi 'I1'
+
+
550
Ef
(11.211)
Ef'
where 110') occupied or 1/'0') unoccupied refer respectively to whclheror not lin) is occupied or [1'(7) is unoccupied in the ground Siale. For S =~, M = the states 11'0') and lIn) can be summarized by a diagram in which solid arrows designate states occupied by electrons, an arrow pointing up. T. referring to u = and so on; and dashed arrows. l or :. referring to unoccupied states. The ground state is shown in Fig. t 1.19. An excited state is obtained by transferring an electron from an occupied to a vacant orbital. For a field in the x-direction, the orbital Zeeman term couples only the slate YI + Y2 to ZI + Z2, as can be seen by an argument similar 10 that for deriving (11.162). The states joined in the g-shift are shown on Fig. 11.20. The explicit evaluation of the matrix elements follows the discussion relating to (11.184) and (11.189) giving
I:
+!.
(lqlu(1) II' q)(I' ql~\5, Ilq')
E,
1<1;111";1'
(lul1i(i)
soo
E"
Wq ')(I'qIJU(1)llq') E/-
(11.208)
Efl
where lu includes all values occupied in 105M) and lu' includes all values occupied in lOS M /). As we have remarked. allowing I' q or r (1' of the excited states to include values occupied in either 105M) or 105M') introduces pairs of terms that cancel. 11 is therefore simplest in practice to reimpose conditions A and B so that no superfluous terms arise in the summation. By means of the Wigner-Eckart theorem, it is possible to show that (11.208) implies that all matrix elements (OSMI'HLlgIOSM') can be obtained from a Hamiltonian of the fonn
1tLlg = 2P
)occupied )unoccupied
!.
Therefore we find (OSMI1i"gIOSM') =
I"
(11.207)
SOO
[(lUIU(I)II'O)(l'UI1i~~llu)+c.c·l
I:
=
E
I -
= !.
(OSMI1i"gIOSM)
L:'(lqIU(I)lI'q)(l'ql1i~2ollq') "
=
L Hqaqq,Sq' '=".J',_ ,'=".,.-
(11.209)
-r
~
---t+-t-1
ZI+ Z ,
y.-y, Xl - X,
----'I I
10
-1-+1--
-+-l,---
Z, -
z,
1 }} or Lhe VI: center. We Fig.ll.tll. Ground stille lIS5ume thd the eryst"l fic~d splits the stlltell Zl ± z, slightly rrom YI ±Y1. The lJ(llid arrows indiCllle 1111 0<;eupicd state (, I I); the dashed one (i) is unoccupied
551
Fig. 11.20. The double arrow indicates states joined by the matrix elements of 11.120) for a field in the :I:-direction, The arrows indicate the electron spin quantization, r signifying spin parallel to the static field and 1 signifying spin antiparallel. The dashed arrow 1 state is vacant in the ground state
+i=::!:±:===: :-t1
Yl + Y 2
=
(Z1l 1\1)IYI)( -~IS\I)I-
D
We obtain in this manner = _ 2PH.(-I/2)IS\I)I- 1/2) (o~22~1'Ii"gIO~~) 22 E + E ZI
X I +Xl
Since the spin Zeeman energy
-Hf----
(11.216)
Z2
(11.217)
yl - Y1
(O! ! l1-lszJO! !) is
(O! ! 11-lszI0! !) ::: 2f3H:I:(! IS~l) I!)
(11.218)
being just that of the one unpaired spin, we have
(
11(1)1
O.!. .!. J1-{~gIO.!. .!.) "" 2f3H :I: (Y1 1L:I:(1)1) ZI (ZI 2 2
E YI +Y1
22
I)
'2" '}{SOOZ Y1 - '2" -
E ZI +Z1
where '}{SOO:l: is the one-electron, spin-orbit coupling associated with the xcomponent of spin and is given by
(Zl - ~['H~go:l:IYI -~) "" 2~;c2 [1 zi(E x p):l:yldr] x
(-~IS\I)I-
D
N
(11.214)
i::: I
For equivalent electrons, the (j'S are all equal. If one has Russell-Saunders coupling in the free atom, the lotal angular momentum quantum numbers Land S are good quantum numbers and for matrix elements internal to a given Land S, we have
1-lS0 ::: AL· S
.
=2(1-
E ZI +Z1
(
E YI -Y2
-
>.
E ZI +Z2 ~ E YI -
)
)
(11.219)
Y1
where ..\ is the free CI atom spin-orbit coupling constant. We note 9n > 2. In a similar manner we get gyy :::
(11.213)
where dr indicates integration over the electron position coordinate. The expression of (11.213) involves functions on one atom only, and can therefore be related to the free-atom expression. In fact for a free atom, if one neglects the coupling of one spin to the orbit of another electron, one can write the spin-orbit coupling of N electrons as
1-lso"" L(iLi ·Si
g.. =2(1+
(11.212)
2
(I -
E ZI +Z1
>.
E:I:\
9n : :
) -:1:2
2
(11.220)
It is interesting to note in our calculation that the origin of the positive g-shift rather than the negative one that we would have for a single electron is the spin matrix element (-!ISil)1 of (11.217). The I-!) states come in because we have excited one of the paired spins into an originally unpaired state. We deal with a spin that points opposite to M. We may contrast the situation with one in which only five electrons fill the states, as shown in Fig. 11.21. We shall assume the degeneracy of the states YI +Y2 and XI +X2 is lifted, as shown, and likewise for Y'-Y2 and XI-X2. A field in Ihe x-direction would join states IYI + Y2, and IZJ + Z2,!) and would make 9 < 2 (an "electron" shift). On the other hand, a field in the z-direction joins the nearly degenerate IYl +Y2. and IXI +X2, slates. It would have g > 2 (a "hole"
-!)
!)
-04)
-!)
(11.215)
If N ;: I, clearly A::: (. If N represents a shell that has only one missing electron, A ::: -(. Since ( is always positive, we obtain in this way the faci that holes have negative A. We can utilize the free-atom ('s 10 evaluate (11.213) since it enables us to write
f------552
ZI -
Z2
!o'ig.lI.2t. Filled states when the V center orbitals contain only five electrons
553
shift). The close proximilY of these two states would make l.dyul > 1.d9zzl, and 9"y, of course, is Slj]] 2, since Ihe peninenl matrix elements vanish. It is clear that Ihe 12 states are less Ihan half full, yel the predominant g-shift is that of a "hole". We see, Iherefore, Ihat we must use extreme caution in characlerizing centers as "eleclron" or "hole" centers simply from the g-shift data. We should also comment that we have assumed very simple functions with no overlap between atoms, to compute the matrix elemenls. In general, we would need to make correclions both for overlap and for the possibilily thai the funClions :1:), Yl, ZI> and so on are linear combinations of alomic orbilals, as we did in discussing the hyperfine coupling. However, these corrections do nOI aller Ihe principles, although they do complicate the numerical calculations.
12. Summary
We have considered a variety of effects - line widths, chemical shiflS, Knight shifts, hyperfine splillings - a bewildering array of seemingly special cases. As we look back, we see some effeclS that occur in first-order penumation theory, others thai require a higher order. Since we have discussed the phenomena one by one, it is appropriate to summarize by wriling a single Hamiltonian that includes everything. As we coOlemplale it. we should remind ourselves of the significance of each tenn. We write below Ihe Hamiltonian describing a nucleus interacling with an electron in the presence of a magnelk field Ho. We define the vector potenlials Ao. associated with the field Ho. and An' associated with the field at the electron owing to the nuclear moment (An = P X rlr 3 nonnally). We also define the quanlity
.
,
1f=7~+-Ao 1
(12.1)
C
Then we have Ihe following Hamillonian: 2
n2 'l.J=_Ji I~ 2mv
.1"",,,,,, m~
k;~'lio
+V0+ V~rrl\
.leo'_ """.Iial '.'/' ;~po'.~';,,1 'h. n.w ._". d... ,~ Io•••ad <>l' .U'....... IOi4t ,h.
.1...1t... 01, oth
1.. ".....
.._
.lec,roo .pi. loc",...
.1"",,,,,, .pi._hi' cOIIpli••
, (p + 2mc .oupli••
..o,u
"
• Ao + Ao • p)+ 2mc2 A~
~r
.I.. ,,~. O'bital ",o'io. '0 Ho
couplio. or uolu. m"",.o' '0 .1""00 O'bi"llDo,io~
"'''''''0'
.....pli•• of uol... .. i,b .1.." .....pio momo" for r-IIol..
couplio. or ~uol..r mo",.01 .. i'b .1"'1'<10 .pio mom..1 for""" r-.lo,..
OOIIplio. of auol... qu.d.llpolo """"'0' '0 fi'ld ••odico' duc ,0 .Ioe'......od ........10_...
ouel...
z..1D.~
'~"17
We can add to Ihis the coupling of nuclei with each other and the magnetic coupling of electrons with each other. 554
5S5
Problems
Chapter 2 2.1 Consider an Hennilian operator F which is an explicit function of time. (For example. F = -"'(Ii.I7;Hx cos wi, the inlcruction energy of a spin with an allernating magnetic field in the x-direclion.) Prove that
dF = '!'[1{ F] + of dth'
at
where 8Flat represents an actual derivative of F(O with respect to lime. 2.2 Equation (2.22a) gives an expression for (JJz(t» for a particle of spin Generalize the expression for a spin I.
!.
2.3 A magnet has an inhomogeneous static magnetic field. The fraction of spins df thai experience a magnetic field between H and H + dB is
= df = r'.H)dH
when:
J
r'.H)dH = I
o Assume the inhomogeneity is slight and thai it simply gives a spread in field with no change in the direction. Take the field 10 be in the z direction. Compute lhe magnetization in the x+direclion, perpendicular to the static field, as a funclion of lime. assuming that al t "" 0 the 10lal magnetizalion was Mo, poinling in the x-direction, for the three fonns of P(H): b)
P(H) is a constanl for Ho - 0 < H P(H) <x C'p ( - (H - Ho)'la'),
c)
p(H)cx. t+(ll
a)
< Ho +0 and is zero for all OIher fields.
~lo)17al'
2.4 A nucleus of spin! is quantized by a field Ho in the z-direction. It is in the m = slate al t = 0 when a rotating magnetic field of amplitude HI is applied for a time two producing a 90° pulse.
+!
a)
Compute lhe wave function of the spin in the rotating reference syslem as a function of lime during and after the pulse. 557
b) c)
Compute the wave function in the laboratory reference system during and after the pulse. Compute the (It;r,(t)) during and after the pulse.
2.5 A coil of length I, cross-sectional area A. and n turns is wound in the fann of a solenoid. The axis of the coil is in the x-direction, and a static field Ho is in the z-direction.
a)
b)
Assuming the nuclear moments are in thennal equilibrium. what is the nuclear magnetization per unit volume Mo prochx:ed by Ho. in tenns of H 0 and the static nuclear susceptibility xo'1 Compute a numerical value of xo for protons in water at room temperature. using the fonnula N'"(2h 2 I(1 + 1)
xo =
c) d)
3kT
where N is the number of spins per unit volume. (For procons. '"( can be found from the fact that the resonance occurs at 42MHz for Ho = 104 Gauss.) The magnetization Mo is turned by a 90 0 pulse. Derive an expression for amplitude Vo of the voltage induced in the coil by the precessing Mo. Make a numerical estimate of Vo for procons, assuming a Io-tum coil 2cm long, t em in diameter. and an Ho of S<XX>Gauss.
b)
show that the solution of these equations is a(t)= ~[a(O)(1 +cos wlt)-c(O)(l -cos WIt) + ihb{O) sin Wit] 1I(t) = 11(0) cos WI t
c(t) ::::
c)
a) b)
Derive an expression for the voltage across the condenser in lenns of the induced voltage Vo of Problem 2.5. Lo, Ro. C. and Q(= Low/fl). Using the numerical estimate of Problem 2.S(d), and assuming the coil has a Q of 100, compute the size of the voltage across C.
2.7 Rotation behal/ior of a spin J particle: Consider a particle of spin I. with eigenstates of It given as "t. "0, U_1 corresponding to the eigenvalues of 1, O. and -I respectively of It. Consider that it is acted on by a rotating field, HI> exactly tuned to resonance as in the discussion of the spinor behavior of a spin ~ particle in Sect. 2.6.
a)
Taking the wave function in the rotating frame, '1//. to be "/(t) = a(t)u1 + b(t)uo + c(t)tt_1
and defining tjons
WI
as ,HI> show that a. b. and c obey the differential equadb .Wl - = I-(a+ c)
dt
5S8
J2
I
.
Sill WI t
+ cos WI t) - a(O)(1 - cos
WI
t) + i hb(O) sin
WI
t]
Show that these equations give "'(t) = tp(O)
if WI t ::::: 2:1f
in contrast to a spin ~ particle for which "'(t)
= -!/J(O)
if Wit
= 2:1f.
2.8 In Sect.2.9 the existence of spin echoes is derived assuming that the "7rn." pulse gives a :lfn. rotation which is positive in the right-handed sense. This implies a negative '"(. Carry through the derivation assuming the :lfn. rotation is positive in the left-handed sense, as with a positive '"(. showing that the echo is formed along the negative y-axis of Fig. 2.13. 2.9 Consider an experiment in which inirially the magnetization is at its thennal equilibrium value. kMo• along a slatic field kHo. At t = 0 the magnetization is invened by a 1f pulse. It then grows back towards its thennal equilibrium exponentially according to the equation
dM z
2.6 Suppose the coil of Problem 2.5 has inductance Lo and resistance Ro and is in series resonance with a condenser C.
4[c(O)(
i
+ ../i(a(O) + c(O)]
-;u=
Mo-M z T1
At a time Ti> M t is inspected by applying a 7rn. pulse and observing the initial amplitude of the free induction signal generated in an rf coil oriented transverse (o Ho. a)
Show that M z • the value of the transverse magnetization immectiately after the :lfn. pulse. is
Mz = MoO - 2e- TJ / TI ) b)
Assume now that owing to natural line breadths, the transverse magnetization decays rapidly to zero. Suppose one now waits a time 1'2. long compared to the time for transverse magnetization to decay. then again applies a 7l' pulse, and again applies a second 7l'n. pulse at time 1'1 later. Suppose one repeats this pulse sequence many times:
... !I+-1'I_1I +-1'2"'" 11+-1'\"'" II··· etc. 11" 11"/2 :If 11"/2 Show that the initial value of the free induction decay following the 11"/2 pulse becomes
M z =Mo[1-2e- T!/'r1 +e-(TI+fl)/'I'I] 559
2.10 This problem concerns lhe calculation of the time at which a spin echo occurs for a (7r(2, 11') echo sequence, when the duration of the pulse is not taken as zero. Refer to the end of Sec1.2.9 and Figs. 2.15 and 2.16 for the notation. The goal is to derive lhe relationship of (2.73), r' = r + IhH I, in the approxi· mation that we neglect corrections to r ' involving (I/H t )2. In all that follows, this approximation is assumed. Thus tan (lJH/H 1) = eH/H t etc. Prove that the "11'(2 pulse" rotates the magnetization, eM, from its initial direction along the z-axis to the plane (where the y and axes coincide), and that it makes an angle LJ.8 = HI/eH with the negative axis. Prove that the projection of eM in the x·y plane makes the same angle. LJ.8. with the y-axis. Show that the effect of the ":If pulse" on the projection of eM in the x-y plane is a reflection in lhe x-z plane, as in lhe "infinite HI" 11" pulse (i.e. there is no correction for lhe ":If pulse" similar to the correction for the ",..(2 pulse''). Deri'e (2.73)
a)
x'_y
b) c)
d)
y
y
(2.73) Show that (2.73) is also valid for an echo produced with a (311"/2. 211") pulse sequence, but that (2.74) becomes
e)
, T
2 =T+3:1ft3r/2
(2.74)
2.11 Consider an operator R and its inverse R- t (Le. RR-I = R-tR = I). An example might be
R
= e i81•
which is associated with rotation about the z-axis. Let G be some operator function which in general does not commute with R (i.e. RG - GR = 0). Prove that R-le iG R
b)
Show thaI the expectation value of the total x-component of spin, obtained by adding up the contributions of all the spins (Le. integrating over dho), vanishes at t = 2r. Show that the time development of expectation value of the total ycomponent of spin (Iy,todm after the 90° pulse (but before the 1800 pulse) is the same as (Iy,tot(t - 2T). the behavior after the peak of the spin echo.
2.13 From (2.170) we have thaI
J "( ')~.!
00
2 I xo=-• wo
w uw
X
°
Show that if X~ax is the maximum value of X" in an absorption line. II
:If
Xmax =
WQ
"2 xo LJ.w
where LJ.w is a suitably defined line breadth. Assuming that the line width of the nuclear resonaoce of protons in water is 0.1 Gauss broad (because of magnet inhomogeneity) and that Ho = 104 Gauss. compute x'~ax for water. and also compute the maximum fractional change in coil resistance for a coil of Q = 100.
2.14 1be response of a cenain piece of material to a step of magnetic field of unit height. applied at t = 0, is M 5tep(t) = xo(l - e- t / T ) a) b)
Compute x'(w) and Xll(w). Show that x' and X" satisfy the Kramers·K.ronig relations.
2.15 In Appendix F. an expression is developed for the complex magnetization MI;+iM y "i,c:H"I"M=oT"["2"+c-T ,.:(",a"'..,+,.:a,,,,,,)I2,,,J MI;+I.M y =-
= ei(R-IGR)
2.12 This problem is concerned with the quantum mechanical treatment of the free induction decay and the spin echo from noninteracting spins subjected to a distribution of static field strength. Consider the time development of the wave function in the rotating frame for a spin echo in which the HI is applied along the x-axis in the rOlating frame. The first pulse is a 1r{l pulse, and the second, applied time T later, is a 1r pulse. Take the distribution of magnetic fields H to be symmetric about wh, so that if ho =: H - who lhe distribution function P(hO) of ho is symmetrical about ho = O. 560
a)
(I
a)
+ O"ar)(1 + O"bT)
1
Show that in the limit of very slow jumping (rew >- 1) the absorption line has the shape of two distinct resonances at frequencies
Ow
w='Y H o±2' b)
Show that in the limit of very fast jumping (rew« 1) a single resonance results at frequency w = '1HO
561
2.16 In Appendix F the Bloch equalions are employed to analyze the case of a group of nuclei which spend equal times on the average in either of two sites at which the resonant frequency differs. This equal time aspect appears in the statement that the two quantities C 1 and C2 are equal, and in the use of Mo/2 in the term involving HI. Suppose the equations were to be used to describe a problem in which the nucleus could jump between twO sites which were not equally populated on the average. For example site "a" might be at a higher energy than site "b" so that the thermal equilibrium populations differed. a} b)
If M a and IIifb are the thermal equilibrium static magnetizations, what must the relationship be of G 1 to C2? Set up the appropriate form of (F.13) for this case and derive the expression for the resulting M~ + iMlJ'
2.17 A static magnetic field H(r} may be expressed as
H = "''' where, is the magnetic potential which, in free space, satisfies
",',,=0 _ If the magnetic field of a laboratory magnet is assumed axially symmctric about the z·axis, show that to lowest order the inhomogeneity can be expressed as
H-HO=Z(O::)
b)
c)
the frame which rotates at "(Ho. Draw a picture of M a and Mb in the x·y plane at a time at which M a makes an angle 9 with the x-direction. Draw a picture of the vectors oMa '" G.Mbot and oMb "" -GtMbot . which would occur if G I is now switched on at the time corresponding to pan (a). Show by vector addition the new vectors M~ "" M a +oMa
Mf, = Mb+oNh
and
Chapter 3
i
3.1 A pair of identical spins of II "" 12 = is coupled by their magnetic dipole moments. Assuming zero external static magnetic field. show that the proper eigenstates of the spins are the singlet and triplet states, and then find the energies of the different states. 3.2 Suppose in Problem 3.1 that a static magnetic field Ho is applied parallel to the intemuclear axis. a) b)
Find the energy levels and eigenfunctions as a function of Ho. An altemating magnetic field is applied perpendicular to the internuclear axis. Find the allowed transi!ions, their frequencies and relative intensities: (1) for Ho much less than the dipolar coupling; (2) for Ho much larger than the dipolar coupling.
where H is the magnitude of H, and Ho is the magnitude of H at the origin and where (8HI8z) is evaluated at the origin.
3.3 Equation (3.43) involves Tr {(1i~, l1-JIJ+}. Prove that this trace vanishes.
2.18 Show, following the methods of Problem 2.17, that the most general form of H(x, y, z) based on including the spherical harmonics of 1"" 2 or less in the magnetic potential is
3.4 Consider two identical spin nuclei. Let I: =: II: + I2: be the total z· component of angular momentum. Evalume Tr {Ii} by explicit evaluation of the diagonal matrix elements for two schemes of quantization: (a) the mi. m2 scheme and (b) the I, M scheme. Show that the answers agree with each other.
H(x.y,z)
=Ho +Ax+By+ Cz
i
3.5 Consider a nucleus of spin I. Compute Tr {I:I7;.} and Tr {I2 Ii}. 2.19 Utilize the expression for H(x,y,z) of Problem 2.18 and the methods of Appendix G to get a more general expression for the effect of diffusion on the echo envelope decay than that of (6.15). 2.20 Appendix F uses the Bloch equations to discuss the problem of a spin which can precess at either of two natural frequencies, depending on whether the spin is in site "a" or site "b". a)
562
Suppose that T2 '" 00, H t "" 0 (free precession), that the quantities CI and C2 are zero, and that at t "" O. M a and Mb lie along the x-axis in
3.6 Consider a group of N nonintcracting spins of spin I and gyromagnetic ratio 1'. The total wave function can be taken as a product of the individual spin states, and the total energy as the sum of the individual eigencnergies. Evaluate the expression for Xl/(w) of (2.190) to give the absorption. For simplicity, let Z "" (2I + I)N and exp(-EalkT) "" I, the expressions for the high-temperature limit. 3.7 The electrostatic exchange coupling between two electrons can be rep· resented by adding a term A5,· 52 to the Hamiltonian. 563
Prove that this lerm commutes with the Zeeman energy.
a) b)
Prove that addition of such a term to the direct dipole coupling does not affect the second moment computed when assuming dipolar coupling
4.2 Calculate the numerical size of the magnetic field produced by the Ofbital motion of an electron in the n = 2, I = I, m = + I state at the nucleus of a hydrogen atom. (Neglect all effects associated with the electron spin.)
alone. 4.3 Consider the states xf(r), yJ(r), and zJ(1') split by a crysral potential A(x 2 - y2). Let the syslem be slarted al time t "" 0 in a state
3.8 Consider three operators, A, B, and C. Prove that
n{ABC} =n{CAB} =n(BCA}
,,(0) = (x+iy) !(,)
,fi 3.9
Consider a nucleus with spin ~ whose Hamiltonian is 1i =
Hz + 1tQ.
where
Hz = 'YnhHoI~ The fonn of 1tQ is similar to the one that sometimes arises when a nudeus has an electrical quadrupole moment. An alternating field is applied to the system to produce absorption. a) b)
Prove that 'Hz and 'HQ commute. Trealing 1iQ as analogous to 1i~ in Seclion 3.3, and assuming A <:: "'(nnBO prove that
(w) = "(nHO and find (Ll.w 2 ).
Compute the expectation value of the z-component of angular momentum, showing that it oscillates in time between the values + I and -1 at the angular frequency of oscillation LJlh. Note that your result corresponds to the classical picture that the effect of the crystal field is to cause the plane of the circular orbit to tum in such a way as to reverse the sense of circular motion periodically. 4,4 The hyperfine coupling of s-states may be found simply in the approximation that a nucleus is a unifonnly magnetized sphere. In this problem, we derive the famous s-state fonnula for that model. A unifonnly magnetized sphere of magnetization M per unit volume can be represented by current distribution flowing on the surface, the current being proportional to M· n. where n is the unit outer nonnal. We shall represent a nucleus by such a sphere of radius R. Consider the current to flow in circles about the z-axis; the surface current density J(8) is then
3.10 Derive (3.61) for the case of a unifonnly TOtating pair of nuclei.
J(U) "" Jo sin 8
Chapter 4
•
4.1 In Section 4.4, gauge transformations are discussed. a) b)
Using (4.18), show that (4.19) is true. Prove that the operator rx
c)
(~V-~A)
for the angular momentum is gauge invariant. Consider an s-state !/J(r) = u(l') in the absence of a magnetic field. A unifonn magnetic field with vector potential Ao = iBo xr is applied. Derive an expression for the resulting angular-momentum expectation value, and evaluate the answer in units of Ii. for the ground Slate of hydrogen, assuming Ho = IO,OOOGauss.
. ""
JoutoC
a) b)
Show that the magnetic field H inside the sphere is unifonn and that outside the sphere it is a pure dipole field. Show that the field inside is
8.
H""T Jok c)
564
J into paper
Show that the magnetic moment 11 of the sphere is (41l'13)R J Jo. 565
d)
Show mal
-H, = JH,lu
2
(')ldT =
8",
2
3"lu(O)1 "
5.2 Consider a system of N spins that interact with one another via a dipoledipole coupling and with an external static field H in the z-direction. Assuming a density matrix f! given by
where u(r) is a spherically symmclric function that does no! vary too rapidly within R of the origin and where H; is the z·component of field
e-'H.jkT
e=
due to the sphere.
4.5 An atom has a single valence electron in an s-Slate and a nucleus of spin I. The electron spin·lallice relaxation time is so short thai the nucleus experiences only the time-average magnetic field of the electron. Derive an expression for the resonance frequency of the nucleus when a sialic field Ho is applied. giving your answer in tenns of the electron susceptibility Xc. Discuss the temperalUre and field dependence: (a) at high temperature (where kT >- "tchHo) and (b) at low temperature (where "fcltRO "" kT).
4.6 In the text the Knight shirl was calculated by firsl-order penurbalion theory. using the fact that the static magnetic field Ho causes a repopulation of the electrons among their spin states. It is possible to obtain an expression for the Knight shirt by using second-order perturbation theory and by assuming that the applied field varies spatially in such a manner that there is no repopulation. Thus, suppose the applied field is in the z-direction, varying with x as H~
=Ho cos qx
Consider a nucleus at x '" O. Using second-order perturbation, show that the electron wave function is perturbed in such a manner that a Knight shift is produced and that, in the limit of q ~ 0, the answer agrees with the result found in the text. 4.7 The elecrronic structure of the hydrogen molecule can be described in terms of the molecular orbilal model, using molecular orbitals that are a linear combination of atomic orbitals. l The lowest molecular orbital is the bonding one formed by a linear combination of free-hydrogen Is-states. Compute an expression for the indirect coupling of the proton spins. As an approximation consider that the only excited state is the antibonding orbital formed from a linear com* bination of the free*hydrogen Is-states.
Chapter 5 5.1 From (5.39) and (5,40) verify that W rnn '" W nrn exp [(Em -En)PLl. where PL '" 1/kTL and T L is the lattice temperature. I See, for example, II. Eyring, J. Waller, G.E. Kimball: QuantlUft CMmislry (Wiley, New York 1944) Chaps. XI and XII.
566
Z
corresponding to thermal equilibrium (where Z is the partition function), show that the thennal equilibrium expectation value of the total magnetization is
(M,) = (M.) = 0
•
(M,) =
N~'!fi32:~ + 1) H
in the high-temperature approximation. It is interesting to nOle that these equations are of the fonn M = CHIT Curie's law, and that the constant C does not depend on whether H is large small compared with the local field due to neighboring dipoles - in contrast to one's naIve expectation.
0;
5.3 .Consider in a metal a system of nuclear spins that interact with a dipolar couphng only. By means of (5,49), prove that the spin-lattice relaxation time in zero static field is one-half its value in a strong field (Tl '" l/aoo).
i
5.4 A nucleus of spin has a static Hamiltonian 'Ho '" -'YnhHoI~. It is acted on by a time-dependent interaction 'Ht(t), given by
'1M')
=A(l)(I; -
I;)
where A(t) is a random function of time. Assume that the correlation function A(t) is
A(t)A(t + T) = A(l)2 e +l/", a) b)
Express 111(t) in tenns of the raising and lower operators, 1+ and 1-. Compute the probability per second of transitions from the m '" state to the other three m-states induced by 11 1(t).
i
5.5 Consider Problem 5.4. Assuming that the relative populations of the mstales always correspond to a spin temperature, compute the spin-Ianice relaxation lime due to the 'H1(t). 5.6 In Section 5.13, the effect of an alternating field is included in the density matrix fonnalism. a) b)
Show that the solutions of (5.398) are correct for low V. Carry OUI the solution for (Mz(t». assuming large V, thereby obtaining the results for saturation.
567
5.7 Consider a system with three energy levels 1, 2, and 3. An alternating interaction V(t) '" V cos wt is applied nearly at resonance with the rransition between states I and 2. a)
Write down the differential equations for the density matrix analogous to (5.394) and (5.395). In the limit of negligible saturation, compute (M:.:(t) and show that the width of the resonance is affected by the relaxation to level 3. (This is the phenomenon of lifetime broadening due to transilions to a level that is not directly involved in the spectral line.)
b)
t/>=
~c"tlnexp( -~Ent)
where en are independent of time, and Un are the eigenfunctions of the timeindependent Schrooinger equation of energy eigenvalue E". Prove that the expectation value of the energy for such a function is inde~ pendent of time, Le., that the energy of the system is conserved. 6.6 Derive (6.46) W(w) =
i"{2 Hfg(w)
from the fonnulas given in Sect. 2.12, using the high temperature approximation.
Chapter 6 6.1 Consider (6.3). Let M:.: '" Mo, My::; M~ ::; 0 at t::; O. Show that for times
6.7 In Sect. 6.4, formulas are developed which show that for slow changes of the magnetic field for which the applied field Hi is always much larger than the local field,
less than approximately TI, M~ and My will remain small, so that M will lie along HI, and will decay to zero exponemiaJly in T2. 6.2 Show that the expression for the average energy given in Sect. 6.3
E::; LPnEn n
can be found as a derivative of the partition function Z with respect to {J ({J == l/kT) thus confinning (6.1). 6.3 Show that the average energy E of a syStem of N nuclei acted on by an applied field H 0 and coupled together through a dipolar Hamiltonian Jid is -
E::; -C
(H6+ H
2
8
r)'
I
(1
and
Mf=Mi
Show that these results can also be obtained by considering what happens to the population of the 2I + I Zeeman levels of the individual spins for an assembly of N noninteracting spins. (The lack of imeraction corresponds to setting
HL = 0.) 6.8 Consider an adiabatic demagnetization experiment similar to that of Sect. 6.6. Show that the sign of the spin temperature in the rotating frame depends on whether HI is switched on when Ho is above or below resonance. For the case of negative spin temperature, draw a figure showing M and He in the :l;-Z plane for several values of He as one approaches resonance.
where 2
CHL = k(2I + 1)N Tr {1td } and 6.4 The entropy
H Hi - r- Or - OJ
c::;
N"{2h 2 I(l + I)
3k
is given by
E+k91nZ 8 Evaluate this expression in the high temperature approximation for a system of N nuclei to show that
6.9 Let HL = 3Gauss, Make a graph of MIMo versus ko for three cases i) ii) iii)
HI = 1 Gauss HI = 3 Gauss HI = 9 Gauss.
Assume that Hl is switched on when ko» Ho in each case.
u=
C (H 2 +H2)
u::;Nkln(2I+1)-2
0
82
L
6.5 Consider a Hamiltonian Ji which is independent of time. We know, then, that the most general solution is
568
Chapter 7 7.1 Show that (7.59), the exact expression for of a forbidden transition, is equivalent to
(I~)
corresponding to saturation
(I,) = ~ exp(A/2kTl [exp (,"'Ho/kTl - exp(-,,,,Ho/kTl] 2 2 + exp(A/2kTl [exp<,"'HolkT) + exp (-,"'Ho/kT] 569
7.2 Show thai the nuclear polarizalion {l:} produced by the scheme of Fig. 7.8 is the negative of that produced by the scheme of Fig. 7.9 if 1A/21 «'eliHo·
a)
(H(mi »eff = i(H,)s
7.3 Consider the nuclear Overhauser effect for a pair of spin! nuclei as in Sect. 7.5. a)
b)
Show that if (WITcP« 1 and (WSTCP« 1 in (7.47), that (7.61) becomes
.'!.«I,) - Io) = '«([0 - (I,» + ~(50 - (5,»
c)
50) =
S(
1Uo - (I,»
d)
+ (50 - (5,»))
where ~ = A~Tclli2. Show that if the initial conditions on (Iz) - Jo and (S:) - So are
(I,(O» - Io = 0 (I,(I» - Io
=
,
(SAO) - So = Sj
~5;["p (-3"/2) - exp (-"/2)]
e)
=!)
a)
Set up the equations analogous to (7.13) for the case of a steady-state solution. Solve the equations, expressing the populations 1'2 and P3. What relationship between Wt 2, W 2 3, and the energy level spacing is necessary to achieve a population inversion between states 2 and 3? What is the corresponding relationship for a population inversion between states 1 and 2?
7.5 In Sect. 7.13, the double resonance method due to Hahn is discussed. Show that the operator T given in (7.91) transforms the Hamiltonian of (7.90) into the Hamiltonian of (7.94) of the double rotating frame. 7.6 Decoup/ing. Consider the spin-Hamiltonian of (7.190) which describes the Royden-Bloom-Shoolery-B1och method of decoupling:
'H' = -''/I;IHoIz -1" s li[(h o)sSz +(H,)sSzl + AIzS z
(7.190)
Since [[z, 'H'] = 0, the eigenstates may be characterized by mI, the eigenvalues of I:. 570
Explain in terms of the figure of part (b) why there are transitions of the [-spins in which ms' may change. lliese are called panially forbidden transitions. For the numerical case of pan (b), find the frequencies of the allowed and partially forbidden transitions of the [-spins.
7.7 Consider a spin echo double resonance experiment in which the echo of the produce a magnetic field [-spins is monitored, and in which the S-SpillS (S ± hS I on the I-spins. Let the I-spins be observed with a 1r12-rr pulse sequence, and the S-spins be flipped with a 1r pulse.
and
7.4 Consider a system to be used to make a three-level maser with energies El> E2' and Ek (Et the highest; EJ the lowest) (see Fig. 7.11). Suppose one saturates the transition between levels I and 3.
b) c)
-!.
that
(5,(t» - 50 = ~5;[exp(-3't/2)+exp(-,,/2)J
a)
Draw a scale picture of (H(mJ»eff for the case A = 1"S li(h o>S, (ho)s = (Hds .for mJ = ! and mi = Show that the energy eigenvalues of (7.190) are given by
x {[(ho)s ~(AmdYsh)12 + (Hd~}'/2
aod
b)
+ k[(ho)s - (ArnII"Ysh»)
E(mJ, ms') = -1"JhHomJ -,stuns'
dl
.'dt!.( (5,) -
Show that for a given mI, the S-spins see an effective magnetic field (H(mI»eff given by
Show that the I echo, M/(2r), varies with
l'
according to the equation
M/(2r) = Mo cos (21"/h/S1')
7.8 Consider the energy levels shown in Fig. 7.4. Suppose one uses an adiabatic passage at the electron frequency (E 1 - E2)/;1 to interchange the populations of levels I and 2 and then quickly observes nuclear resonance. a)
b)
Show that the nuclear resonance transition (E2 - E 4 )/h has an increased absorption rate and compute the ratio of the rate of energy absorption to its normal value. Show that the nuclear resonance transition (E: ~ Ea)/li has a stimulated emission, and compute tre ratio of the rate of energy emission to the nonnal rate of energy absorption of this transition.
7.9 The density matrix of a pair of spins [ and S in thermal equilibrium at temperature 8 is, in the high temperature approximation, ) 11<8) = Z1 ( 1+ hHo k8 <11 1, + 15 5 ,) Consider the case of an X/(rr!2) pulse. Then, the time-dependent expectation value of J+ is
571
(I+(t)) = T' (r+ e(l» := Tr {/+ exp(i(fll f: + flsS:: - alz Sz)t)X,e(8) x X,I exp (-i(n!I: + nss: - aI:Sz)t)} Show thai the term in (I+(t» arising from Sf in f!(8) can be wriuen as (1ShHoIZkS) x Tr{I+S.:}. which vanishes. 7.10 Using the relationships for spin
'i panicles:
c iS• S = cos (5:9) + i sin (5:9)
will not produce an echo for a system of two identical spin by a Hamiltonian
! particles coupled
1t.:::: AlrSr 8.4 In order to grasp the meaning of the various spin-flip narrowing pulse sequences. one can begin by assuming the dipolar broadening is negligible. that there is negligible magnetic field inhomogeneity. and follow what happens to the magnetization. Employ this method to the three-pulse cycle of Sect. 8.12 to verify the results of Fig. 8.14. 8.5 Employ the method of Problem 8.4 to describe the magnetization vector in the rotating frame through the first cycle for the four-pulse sequence (r.X. T. Y.2T. Y.r.X. r).
cos (5:9) = cos (On) sin (5::8) = 2S: sin (8/2)
Chapter 10 Show thal
e iS• SSze- iS• 8 = Sz cos () - Sy sin 8
10.1 It is stated in Sec\. 10.3 that (lO.27b) and (10.31) provide a set of recursion relations among the elements (J MJI'/ITLM IJ' MJ''I1') for the various possible values of MJ. M'. and M J ,. and a fixed sel J. L. J'. 1'/. '11'. For Ihe case that J :::: J' show Ihal it is indeed true that specifying one matrix element (for example, that for which MJ :::: M J , :::: J) enables all others 10 be computed by using Ihe recursion relations.
Chapter 8 8.1 Suppose that a spin echo is produced by applying the H I of the first pulse along the +::z:-axis in the rotating frame, rotating the magnetization to the -y-axis, and a 7( pulse a time T later with H I along the +y-axis.
10.2 Verify that the functions Tu,.,(J) of Table to.l in Secl.IO.3 satisfy the commutation relations of a T 2A1 with respect to J.
Show thaI an echo is formed at 2r along the -y·axis. Show that if HI. when on, is produced by applying a linearly polarized allernaling field in the sample coil along the laboratory frame x-axis
10.3 Consider an axially symmetric potential and a weak static field The Hamiltonian is then
a) b)
Hx{t)=2HI
coswt • t
11.
,
the situation described above would be produced by changing Hx(t) to Hx(t)::::2H 1 cos(wt-i)
for
t>r
8.2 Draw vector diagrams for the TOtating frame to prove that Fig.8.1 correctly shows the echoes in a Carr·Purcell pulse sequence. 8.3 Apply the vector model and an argument similar to that in (8.12-20) to show that a pulse sequence X{7r{l.) ... To 572
••
X(7r) ...
2
e qQ 3 2 2 :::: 41(2/ _ I) ( I: - I ) -
'Yn liH l r ,
If H :::: O. the spins are quantized by the quadrupole coupling as shown in Fig. 10.4. The states 111 :::: ± are degenerate. Show that when H is weak. these states are split, Ihe energy difference going from '"fn1i.H when z' is parallel 10 Z 10 (I + !)"YNliH when z' is perpendicular to z.
!
10.4 Consider the Hamiltonian and energies given by (10.65) and (10.66). An alternating field Hz cos wt is applied perpendicular to the z-axis. Find the allowed transitions, their frequencies. and Ihe relative intensities. Work oul numerical answers for the cases 1:::: and
¥
!.
T •..
573
Show that a charge e located at a point gradient 10.5
Vzz
==
8'V)
of
--2
8z
Z,II,Z=O
Vzz
xo.
YO. Zo produces a field
11.3 Equation (11.65) says
_ 3zij - rfi
- e
"!-IS = 'Ye'Yn h
5
ro
'( T)
2 z 5 z - I· S} r 3 '5(3l
Show that this follows from (11.64):
'J r
10.6 A nucleus of spin ~ experiences an electrical quadrupole coupling
'HIS = 'Ye'Yn h
1
I [ (I·r)(S·r) 2 , r2 - 1·5 x f (r)dT 3 3
'H = AU; - I;) a) b)
Show that the energy eigenvalues are 0, ± 2/7 A. Show that the eigenfunctions are
_
11.4 In Sect. I 1.3 the hyperfine splitting was worked out for the case of an isotropic electron g-factor. Generalize the result to the case of a Hamiltonian for a field with x- and z-components only:
{~I = (f.)I/'[~'/2 - (W/2~_3/']
E-O
~2=(-&)1/2[~_'I,-(~)1/2~3/2] "" =
E - 2(7)1/2 A -
{
m1/2~_3/21
~4 = (fa) 1/2 [~-'/2 + (¥) I/'~_I/' +
~, = (fa) 1/2 [~'I' - (¥) 1/'~I/' +
E=_2(7)I/'A {
c)
(fa) 1/2 [~'/2 + (¥) 1/2~1/2 +
1-{ =
m1/2~3/,l
m1/'~_3/'] ]
~6 = (fa)1/2[O_'/2 - (¥)I/'~-1/2+ m1/2~3/2
. where ~5/2 is an eigenfunction of I z .with m = ~.' and SOl o~.. Show that application of a small static field Ho In the z -dlI'CCtlon spilts the degenerate E = 0 states, the splilling being 'YnhHo(¥-) independent of the orientation of z' with respect to x, y, or z.
P(gzzHz + gzzH z ) + A z l z 5 z + A y l y 5 y + A z I,5, - 'Ynh(Hzlz + H::I,) .
11.5 Consider matrix elements (O'ILa,lz). where O',er' = x.y.z, which are frequently encountered in problems for which Ix) is the ground electronic state. Show that there are only two which are nonzero (0' = y, 0'1 = z and 0' '" z, 0" = y) and evaluate the matrix elements for them. These results are useful for the following problems. 11.6 Consider an atom containing a single p-electron acted on by a crystal field giving an extra potential energy A(x 2 - y2). The resultant energy levels are shown in the figure (neglect spin ). Assume L!» kT. Compute the principal values of the temperature independent susceptibility tensor.
10.7 Prove that the eigenvalues EI, F:2, and so on of a Hamiltonian 'H = AU; - /f.) come in pairs ± £1> ± B2, or else are zero. (Hint: Consider the
effect of :n operator R that changes x into y and y into -x).
Chapter 11 11.1 Evaluate the coefficient GJ> defined by (11.41), to obtain an answer analogous to (1 1.53a).
In the notalion of Secl.11.2 prove that, with respect to L z , L y • and Lz, the function x 2 - y2 is a linear combination of TU.f's.
11.2
574
11.7 Consider the atom of Problem I I.6, but now include the spin. Let there be a spin-orbit interaction ),(r)L· 5 added to the Hamiltonian. Take as basis states the states IO'ms) where 0' = :t.y, Z and ms = ±~' the spin quantum number.
a)
Show that the matrix elements of the spin-orbit coupling (ul.\(r)L . SI:ts') which are inlemal to the ground orbital state vanish. 575
b)
cJ
Using penurbation theory, show that the spin-orbit coupling mixes ex· cited states into the original ground state wave funclion IX3) and find (he corrected wave functions Ixs)corr correct to lerms linear in A. Assuming an applied magnetic field along the z-direclion, compute the lotal orbital plus spin Zeeman energy by first order penurbation theory, i)
ii) d)
b) c) d) e)
•
•
1{::-2 (PoAn+Anop)
me
(I)
where An :: 'YnhI X r/r 3 is the vector potential due to the nucleus.
using functions Ixs) and using the corrected functions of (b).
a)
Show that the Zeeman splitting of (c)-(i) is L1E = 2{3H but for (c)-(ii) it is of the fonn L1E =: gf3H, and find the expression for 9 correct to terms linear in A,
11.8 An atom containing a single p-electron located at the origin is acted on by a set of 6 equal charges, q. The charges on the x- and y-axes are a distance a from the origin, those on the z-axis a distance b. Assume b < a. Neglect spin.
a)
11011 Consider a single p-electron whose spin and orbit are strongly coupled so that the J MJ scheme of quantization applies. The nuclear moment and electron orbital motion are coupled by a Hamiltonian
Using the Wigner-Eckart theorem, show mat for matrix elements diagonal in the electron quantum number, J, (1) is equivalent to an effective Hamiltonian
1ie ff:: AJJo] b)
where AJ is a constant for a given J, independent of M J . Find AJ for me J :: ~ state.
Show thai the lowest nonvanishing term in the Hamiltonian which de· scribes spliuing of the p-states is 1{ = A(3z2 - r 2} and determine the sign of A. Find the proper eigenstates and energies in terms of A and (r 2 ), the mean square radius of me p-state orbit. Suppose a magnetic field is applied along me z-direction; compute the 3 x 3 Hamillonian matrix. Find the eigenstates which resolve the degeneracy. For which states is the angular momentum quenched?
11.9 Consider the system of Problem 11.8. Assume that the splitting of the p-states produced by the crystal field is large compared to kT, but that pH <:: kT. a) b)
Show that when b> a the crystal splilling of the p-states has the opposite sign from its value for b < o. Curie's law states that the magnetic susceptibility tensor per atom Xoo' (0:, a' :: z, y, z) goes inversely with temperature Xero:'::
Cerer' T
Show that for one sign of the crystal field Curie's law applies, show that for that sign the Z-, y., and z-axes are the principal axes, and evaluate the corresponding C zz , C yy , and Cn.
11.10 Consider an atom with a single p-electron, the orbital angular momentum being quenched by a crystalline field such as that of Fig. 11.1. By using second-order perturbation, show that interplay of the spin-orbit coupling )"L S and the coupling (d2mc)(p· An + An p) between the nuclear moment and the electron ocbit give an effective spin-spin coupling between the nucleus and electron. 0
576
0
577
Appendixes
A. A Theorem About Exponential Operators We wish to prove a theorem about the exponential funclion of two operators, A and B. and their commutator C:
C" [A.B]
.
(A.I)
The theorem states that when both A and B commute with C, then
e A+ B
=
e Ae B e- C / 2 or
e A + B = e B e A eC / 2
(A.2)
.
(A.3)
We shall prove (A.2). The problem is most readily solved by considering the function exp [>.(A + E)l. We seek the function G()') such that (A.4) To find G()'), we seek a differential equation in >. which it satisfies. In essence this amounts to finding the way in which the function exp [>.(A + E)] changes for small changes in >., and then integrating from>. :: 10 ). = I. By laking the derivative of both sides of (A.4), we get
a
(A
+ B)e),(A+B)
= e>.A(A
dG + B)e>'8C().) + eAAe>'B_
d!.
(A.5)
By utilizing (A.4) and multiplying from the left by exp(-..\B)exp(-..\A), we can rewrite (A.5) as
e-~Be-~A Bc>'Ae>'BG
_ BG = dG
d!.
(A.6)
We can evaluate the expression e-~A Be>.A
== R(..\)
(A.7)
as follows: by taking the derivative of both sides of (A.6) with respect to ..\, we find e~>'A(BA _ AB )eM = dR
d!.
- C=
(A.8)
579
since AB - BA =: C commutes with A. Integrating (A.g) we have R(>.) = -C>'+constant
t
-
-=r
(A.9)
.
We can evaluate the constant by setting>. '" 0 and by noting from (A.7) that R(O) = B. Thererore R(~l.-C~+B
h we get
J
~/I(w) = _w_ r:.e-E./kT 2kTZ -co a,b
(A. 10)
By substituting (A. 10) into (A.6) and using the fact that C commutes with
B, we get dG
-~CG=
(A. II)
d~
x (al".lb)(bl".la)exp {(ice. - E,ltVhjexp (-iwt)dt
(8.5)
We can use the fact that the states la) and Ib) are eigenfunctions of the Hamiltonian 1i to express the expression more compactly as Xl/(w) ==
which can be interpreted to give
~ 2kTZ
J ~(ale-1t/kTe
+~
~,
i7iL / Ait
-co a,
X (blpzla)e-ilJldt
(A.12)
C '" exp { - (>.2Cfl + const»)
(B.4)
-
•
e- i1tt / Alb)
.
(B.6)
But the summation over a and b is clearly just a trace, so that
The constant must be zero because, from (A.4). C(O) = I. Therefore
e A +B = e Ae Be- C/ 2 . Q.E.D.
(A.13)
X"(w) == 2k;Z
J
Tr {e- 7i / kT ei1it / l itze-illt/A J1r }e-ilolldt
(B.7)
-~
In the high-temperalUre approximation, we replace exp (-'H/kT) by unity. If we then define the operator Pz(t) by
B. Some Further Expressions for the Susceptibility
flz(t) == ei1tI / Altze-i1tt/h
(This appendix requires familiarity with Chaps. 2, 3, and 5.) Equation (2.190) gives an expression for XII. Another expression is frequently encountered in the literature. It provides an alternative derivation for the moments of the shape function. It can be obtained from (2.190).
(B.8)
we can also express (8.7) as +~
XI/(w) == 2k;Z
J
(B.9)
Tr{Pr(t)JJz}e-ilJ1dt
-~
X" = ; ;
Le-E./.T!(al".lbll'6(E. - E, -
hwl
(B.I)
.~
by use of the integral representation of the 6-fuoction:
1 +co
.
~)=hJe-~rh.
(B.~
-~
By substituting into (B.I), we obtain
J~e
xl/(w) 1 jew) = -w- == 2kTZ
+~
hw w =-x"() 2kTZ
t...-
-co E ., E•
+J~
.
Tr {/lz(t)/lz }e-llJ1dt
.
(B.IO)
-~
-E /kT
•
x (alJ1rlb)(bIJlzla)exp [i(Ea - Eb - llw)TJdr
We can prove another interesting theorem by taking Fourier transform of (B.10): ,
(8.3)
and substituting for the variable r a new variable t that has the dimensions of time, 580
The quantity Tr {pz(t)J1z} is a form of correlation function. and (8.9) states that i/(w) is given by the Fourier transform of that correlation function. By using this expression for Xl/(w), it is easy to show that omission of the dipolar terms C. D. E. and F of (3.7) gives one absorption at the Lannor frequency only, but that their inclusion gives absorption at 0 and 2wo. We get also a very compact expression for the shape function jew):
J ..
+~
I
I 2kTZ Tr{l-lr(t)/lr} == 21f
j{w)ellJ dw
(B. II)
-~
581
We see that, setting t t
=0, I
2kTZ T,{",(O)",). 2.
+=
J
(B.12)
f(w)dw
-~
By taking the nth derivative or(B.I1) with respect to t, and evaluating at t = 0, we find
I
1 tr' (1)" 2kTZ dtn Tr {llz(t)llz} t=o = 211"
+~ "
J
(B.13)
w f(w)dJ..J
-~
We get, therefore, a compact expression for the nth moment of the shape function few): _
J
(w") -_
wnf(w)dw
-~
+J~
__ (i)-"(tF/dt")Tr (JlzO)llz}lt T, (",(O)",)
0
(B. 14)
f(w)dw
-~
As an illustration let us derive an expression for the second moment (w 2 ). Taking the derivative of Ilz(t) gives
~Tr(ei'Hj/l" e-i'Ht/l" J dt 2 ,..z ,..z
.G)'
T, {.m'I'[1i. [1i. ",)).m'I' ",}
= - hl2 Tr {ei'Ht/ll (1i'llz)e -i'Ht/lI ('H:,llzl} (B. 15)
Therefore
(w') •
_2.. T, ((1i, ".)'j h2
Tr {Il;}
(B.16)
This formalism provides a very simple way of generating expressions for the higher moments. Note, however, that the odd moments all vanish, since few) is an even function of w. So far, apart from assuming the high-temperature approximation, we have left the specification of the Hamiltonian completely general. We can proceed further if we assume that it consists of the sum of a Zeeman term Hz and a tenn 'H:p , often a perturbation, which commutes with Hz. A typical H p is the terms A and B of the dipolar coupling. Then, since H p and 1{z commute,
Tr {e i('H p/ll)1 /lye -i('H p/ll)t Jlz} = Tr {ei('H~/lI)l(- Jjy')e -i('H~/1t)tJlz' } = _ Tr {ei('H~/")1Jly'e(i/ll)'H~1 /J z '}
Tr {Ilz(t)pzl = cos WOt Tr {ei{'Hp/l)t lJze-(i/l)'HptJjz)
(B.20)
Since this is the Fourier transform of the shape function f(w), we see that the transient behavior consists of a term cos wot multiplied by an envelope function. If we define Il;(t) as (8.21)
p;(t) '" e(i/l)'Hplllze-(i/l)'Hpi
we can say that the envelope function is Tr {Il;(t)pz}. By writing the cos wot as cos wot = !(eiWOI +e- iWOI )
(8.22)
we can say that the two exponentials correspond to lines at +wo and wish to discuss only the line at +wo, f+(w), we can therefore write f+(w) =
4k~Z
-woo If we
+~
J
Tr {p;(t)/Jz }e+ilotole-ilotldt
(8.23)
-~
(B.24)
This can be rewritten as
J
1 +~ 4kTZ
'" e(ilh )11' pte -iwoJztJlze""O / z!e -(i/h )1t pt
(B.t9)
But the last trace is clearly the same as the first. Therefore the trace is equal to ilS own negative and must vanish. We have, then, that the correlation function Tr {llz(t)llz} is given as
_1_ Tr {1l;(t)Il'z} = -2
IlzO) :: eOlh )(1l z +'Hp)tIlze -(ilh)(1tz +'Hp)t
= e+(i!h)'H pt(llz cos wot + Jly sin wot)e-(i/h)'Hpt
In general, if 1ip is invariant under a rotation of 180° about the x- or yaxes (as is usually the case), the second term vanishes. This can be shown by evaluating the trace, using coordinates x' '" x, y' '" -y, Zl = -z, which differ only oy a 1800 rotation about x. Then 'H. p = H~ by our postulate, so that
71'
!+(w)ei(w-loI(J)fdw
(B.25)
-~
(B.17)
By laking derivatives as before, we now get +~
J
where we have used (2.55). We have, then,
(w - wo)f+(w)Jw
Tr {Jlz(t)llz} = cos wot Tr {e(i!h)'Hpt Ilze-(i/")'Hptllz} + sin wot Tr {e(i!h)'H pt Il y e -0/" )1t pt Ilz}
58'
(B.18)
(d"ldt") Tr {[Il;(t)pz]/ oj • Tr {1l:(O)J'z}
(i")=-~~~::--- +~
J
f+(w)dw
=i"{(w-wo)"}' (B.26)
-~
583
This, then, gives the nth moment with respect to the frequency wo . This fonnalism has supressed the line at -we, which would, of course, make an inordinately large contribution to wo)"} were it included! By following steps similar to those of (8.15), we now find
«w -
« _ w We also
wo
),)=-.:.. Tr{[Hp,p.)'} h2
Tr {pi}
(8.21)
see that
(w - wo) =
I
Ii Tr {[Hp, P,Jp,}
(8.28)
which can be shown 10 vanish if 1ip consists of lhe dipolar tenns A and B, as shown in Chapter 3.
C. Derivation of the Correlation Function for a Field That Jumps Randomly Between ± h o We shall assume the field jumps randomly between the two values we shall label as states I and 2. We shall call
+ho = HI
- ho = H2
± ho, which
,
ddP2 = W(PI - P2)
dT
.
(C.S)
T
This is a "nonnal modes" problem, with solutions obtained by adding or subtracting: PI(r)
+ neT) = conSI. (= 1 from normalization)
PI(r) - n(r) = Ce- 2WT
(C.6)
where C = PI(O) - 1'2(0) = PI(O). Since nCO) vanishes and since PI(O) = I, C=1. By making use of (C. I), (C.4), and (C.6), we have H(O)H(T) = HI (HIPI(T)
+ P2(T)H2]
= h~e-2I1'T
(C.?)
An identical answer is found for H(O)H(T) if the field is assumed 10 be H2 at T = O. We must weigh these equally (Ihat is, average the answers over the initial fields) 10 gel Ihe final ensemble average. We denote this by a double bar, 10 indicate the fact that we have averaged over an ensemble of initial conditions as well as a variety of histories for a given initial condition: H(O)H(r) = hije- 2WT = G(T) . This is the correlation time assumed in Chapter 5, with 2W
';;!
llro·
(C. I)
Then we wish to know the correlation function G(r): G(T) = H(t)H(l + T)
dpi = W(P, _ PI)
D. A Theorem from Perturbation Theory (C.2)
where the bar indicates an ensemble average. If the field is HI at time t = 0, then we can write for a single member of the ensemble: (C.3)
where PI(r) and P2(T) are either zero or one, depending on whether at time the field is HI or H2' We now perfonn an ensemble average of (C.3) over the various histories. This replaces quantities PI (r), P 2(r) by their ensemble averages PI(T) and l>:2(r), which are the probabilities that in an ensemble in which the field was HI at r = 0, it will be HI or H 2 at lime T. Thus we have
In this appendix we shall derive from perturbation theory a theorem that has wide utility in magnetic resonance. It is closely related to second-order perturbation theory but gives the results in a fonn particularly useful when there is degeneracy. A typical situation in which the theorem has great use is illustrated by the g-shift calculation of Section 11.2. We may divide (he Hamiltonian into three teoos;
1i = 1£0 + 1£1 + 1i2
where
(D. I)
T
p'
1£o=-+VO+VI
2m
HI = 2pH· S
(D.2)
H, = AL·S+pH·L
(CA)
This equation, of course, assumes thai at T = 0, H(r) = HI, so that as r --+ 0, PI(T) --+ I, n(r) --+0. Equally likely is the situation that the field is H2 at T = 0, which will give a similar equation except thai I and 2 are interchanged. We shall assume the behavior of PI and 1'2 as a function of r 10 be given by a rale equation: 584
Since 1£0 does not depend on spin, its eigenstates may be taken as products of an orbital and a spin function. We denote the orbital quantum numbers by 1 and the spin quantum numbers by a. Then
Hoi/a) = Ed/a)
(D.3)
The states 110') are degenerate for a given 1 because of the spin quantum numbers. 585
The tenn 1{t lifts the spin degeneracy. Since it depends on spin only. it has no matrix elements between different orbital states: (D.4)
In general the matrix elements of 1{l between staies 110) and Ila') where a #a' will be nonzero. Therefore the presence of 1{1 still leaves us a group of submathese trices (lal1{llla') to diagonalize. For our example, since the spin was submatrices are only 2 x 2 and are easily handled. The presence of the term 1{2 spoils things, since 1{2 joins states of different I. However. as a result of the quenching of the orbital angular momentum. the matrix elements of 1{2 that are diagonal in l vanish:
(D.7)
where dr and drs represent integration over spatial and spin variables, respectively.. By utilizing (0.6) and the Heonitian property of 5, we have
J
¢iO/1{¢I,O',dr drs =
!.
I)
~~~
",
I] Il
I,
{---1---, ---~---·~---i--'::
{
,,
,,
:
:
----,--, ,
,__ ---'-----,--, , , I
o
'HllIO/l'Zt'rU
0'H2/l
01l ' Uro
I_::..l:i:_:_::'--i--_-l.i-_--.J-_-_--'--l--_-
.-
'
"
= (IO'le-iS1{eiSll'O")
(0.8a) (0.8b)
where we have used the notation 110') for matrix elements calculated using the 1/J's. We may interpret (0.8) as saying that we can look either for transformed functions, rPIO/. or a transfonned Hamiltonian, (0.8b). If we define ?i' as (D.9)
We may schematize things as shown in Fig. 0.1, where the Hamiltonian matrix is illustrated and where we have labeled which tenns 1{1 or 1{2 have nonvanishing matrix elements. Il
1/;iO/e -iS1{e iS 1/J/IO/ldr drs
(D.5)
(laIH,lla') = 0
I,
J
.
Fig.D.1. Hamiltonian matrix. The regions in which nonvanishing clements of?i l or 1i2
we may state as our goal the detennination of a Hermitian operator 5 that generates a transformed Hamiltonian ?i' such that ?-t has no matrix elements between states of different I. Presumably S must be small, since the original Hamihonian 11. has small matrix elements off-diagonal in 1. Therefore we may approximate by expanding the exponentials in (0.9): ?i' = e-iS?ie iS = (l-iS-
~; + ... )?i(I+iS- ~; + ... )
. S' H - -2HS') = H + .[H, SJ + ( SHS - T = H + i[H, S] - ~[[H, SI, SI
may be found are labeled by the shading. The quantum numbers II, 12 lind b designate different eigenvalues of 1io
Writing ?i =?io +?il + ?i2. we wish to choose S to eliminate 11.2· By writing out (0.10), we have
The technique that we shall describe below in essence provides a transformation which reduces the size of the matrix elements of 11.2 joining states of different 1. In the process, new elements are added which are diagonal in 1. In this way states of different I are. so to speak, uncoupled, and we are once again faced with diagonalizing only the smaller submatrices diagonal in I. The basic technique may be thought of fonnally as follows. The set of basis functions 1/J10/ fonns a complete set, but has the troublesome 11.2 matrix elements between states of different I. We seek a rransfonned set of funclions ¢/O/ given by • _e iS •I• (D.6) '1'/0/ '1'/0/
1{1 :: 1io + 1{t + 11.2 + i[?io +?i h Sj + i[1t2' 5j - ![5. [S,?ill
where S is a Hennitian operator that reduces the size of the troublesome matrix elements. In tenns of the ¢'s, the Hamiltonian matrix elements are
(D. 10)
(D. II)
We can eliminate the third teon on the right by choosing ?i2 + i[1{o + ?iI, S] =
a
(D.12)
Then we have 1{1 = 1{o + 1{1 + i[11.2. 5]
i2
i2
+ 2 [[Ho + HI, S], sl + 2 [[H" 5], S]
(D.13)
If ?i2 were zero, S would vanish. Therefore, we expect 5 will be of order 1{2. and the last teon of order (?i2)J. Neglecting it, and utilizing (0.12), we have i 1t'=Ho+HI+ [H"SJ . (D.14)
Z
586
587
Equmion (0.12) may be put in matrix fonn to obtain an explicit matrix for S. Using the facts that 1i1 has no malTix elements between states of different 1 and that 1i2 has none diagonal in 1, we have
L:
(f0'1 1i211'0'1) + i
(fO'I1-£'jfO'I)
i
=
= E/ona , + (lal 1t llla
+~ L
[(lO'I1io + 1i III"0'")(1/1 alii S II'0'/)
I"a"
.
(0.15) =
0"
-(lO'ISI/' 0''')(1' a"I'H JlI'0'1)] = 0
(0.16)
If I f.l l , we may neglect the tenns in 'HI! as being small compared with those
involving E, - E". Then
=.; (lal1i2 1110") (E" - E,)
(0.17)
1
If 1= 1' , we have, for (0.16) L:(lO'I1ill/all)(la"lSlla') = L(faISlla")(la"IHllla')
(0.18)
0"
0"
This is readily satisfied by choosing (0.18a): ll (laISlla ) = 0 .
~(lcrll'H2' Sill, cr')
=
~ L
L
(10' JH2
I" ,0"
EI-E/"
E. The High Temperature Approximation In severnl places in the text we make use of the high temperature approximation. For example, on page 63 in Chapter 3 we replace the exponentials by unity in the expression for X"(w):
[(lal'H2I t 'a")(I"a//ISll'a')
i'(w) =
- (la ISIlII 0 11 )(11/ a ll l'H211' a')]
~
W' 0'//)(1'/ a" JH2IIa')
and neglecting the coupling between states of different I.
2 I"a"
=
lll EIOaa l + (fal1i IlIa') + L (lal'H',d a")(11/ a"I H 211a') I",a" EI EI" (0.21)
(0.19)
since 'Ho and 'HI are diagonal in 1. Thus,
(lal1i'll'a')
[(laIH21///al/)(I//a Ii ISI/....l )
If 0' = ai, we recognize that the tenns in 'H2 give the familiar expression for the energy shift in second-order penurbation theory. However, our expression also includes matrix elements for 0' f. 0". In this connection we wish to emphasize that in degenerate penurbation theory, ordinarily one must find zero-order functions that have vanishing off-diagonal elements. The method we have described places no such restriction on the basis functions 110'). If the quantum numbers 0' lead to elements (lO'IH'lla') between states of different 0', it means merely that we must still diagonalize the matrix (laIH'lla') of (0.21). We conclude that the presence of a tenn 1-£2 is to a good approximation equivalent to adding to the Hamiltonian Ho + 'H I matrix elements diagonal in I of
(0.18,)
Therefore S does not join states of the same I. [Equation (0.17) and (0.18a) enable one to verify that S is Hennitian; that is, that (l0'151/'al ) = (l'a'1511a)*, where the star indicates a complex conjugate.] By using (0.17) in (0.14), we may find the new matrix elements between 1/0') and 11'a'). First we note that the states off-diagonal in 1 are
(IQI'H'II' cr') =
)
- (falSII" 0'1/)(1// 0"11 'h2 Ila')J
(l0'11i211' a') + i(EI - E/I )(fa lSI t 0'/) + i L [(10' l'h I I/a l/)(1O''' IS\l' 0 ' )
(l0'ISI1'(;/)
l
/1J,a"
-(loISII"0")(1'1 O'"I'Ho + 'HIlt 0")] = 0
Th"
(lcrl'Ho + 'HI + 2 ['H2, SlIlcr')
where
L (laIH21/"a//)(f"a//I'h21/'a')
:~~ L::e- E,,/k7'l(alll x lb)1 2o(Ea ",_
Eb - tiw)
(E.I)
1"&" X
[Ell
~ EI" + EI ~ EIII ]
(0.20)
The off-diagonal matrix elements are therefore reduced in the ratio of 1i2 10 the difference between eigenvalues of 'Ho, and the states of different 1 are "uncoupled". The matrix elements diagonal in 1 are modified, too. They become, using (0.14) aod (0.17), 588
is the partition function. Since the energies E a are energies of the N-panicle system, they may range from -N'YhHo1 to +N'YhHo1 as a result of the Zeeman energy alone. Of course, the energy -N'YhHoI would occur only if all N spins were in the state m = I, and is thus quite unlikely on a statislical basis. However, we expect to find typical values of lEal';::' .../N'YhHoI combining the m values 589
of the N spins at random. Since N ';;t 1023 in a typical sample, how can we approximate Ea/kT <:: I? I! is clear that no one spin inleracts with many others. so that in some sense we do not really need to consider 1023 spins to get a fair precision in computing ;I'. That is, assening that 1023 spins are involved is really a fiction. After all. in an applied field of reasonable strength we can predict the location of the absorption by considering only one spin. We believe that the high temperature approximation will hold if the energy of a sing/I! spin is small compared to kT. We wish now to demonslrate how that comes aooUi. To do so, we shall consider a simplified case. that of N noninteracting identical spins, and assen that a similar argument should hold for interacting spins provided that the effective interaction between pairs is still small compared to kT so that no drastic phenomena such as ferromagnetism results. If the spins are noninteracting, we can choose as exact quantum numbers the individual spin quantum numbers ml.m2, ... ,mN. The energy Eo. then becomes
If we define m =
L
(E.2)
Tnj ::::: -liwoM
where Wo is the Larmer frequency. and M::::: operator 1J.z are
L:j mi' The wave function
"( ) _ 1rhw
X w - kTZ
(E.4)
'"
L.J ml,m1, ... mj-l,mj+I, ...
L
x
j ,m j
i....)
exp (mllwolkT)
exp (-m jhwolkT)J(m j I/t",j Imj)1 2
,mj
x 6[llwo(mj -
mj) -
llw]
(E.9)
Here we have used the fact that the sum over the N - I coordinates omitting
Tni is independent of which j we omit since the spins are identical. But
L
Z =
exp (mliwolkT
,mj-t,"lj+l.
L
exp (mj1iwolkT)
(E. 10)
IIIj
If. now. ltIwo <:: kT, we can replace the exponentials in the Tni sum by unity, and obtain
la) and (E.3)
(E.8)
Then•. using the fact that (ml>m2•...• mi •... I/J"'jlml.m2 •...• m = (mjlp:",jlmjl>. (E.?) becomes
ml,m2, ...
j:::::l
mj. we can write
exp (liwoMlkT) ;;: exp (llwom/kT)exp (tlWomjlkT)
N
Eo. ::::: -'YlIHo
},f -
L
exp(m;1lWo/kT) = (2J + I)
(E.lI)
m; giving
L
Z = (2l + I)
exp(mhwolkT) = (2l + I)Z(N _ I)
"II ,m2, ... ,fIlj-l,m j+ I ...
Then
where ZeN - I) is the panition function of N - I panicles. The sum
l(al"zlb)I'
L
(al 2:>zjlb)(bl LI'z.la)
=
j
•
L(al"zjlbXbl"z.la) .
=
(E.5)
j,k
Since the 1J.zj's involve the coordinates of only one nucleus. we see from (E.3) that we only get tenns in which j = k so that 2 l(al1J.:elb)12 = l(ml,m2 •...• mi' .. ·11J.:ej!m\,m2•... ,mi. ... )1
L
(E.6)
j
giving us
exp(+mhwo/kT)
now factors oul of the numerator of (E.9) and out of Z giving
"()_ 1r"W ~ X w - (21 + l)kT ~
tlW X (w)= kTZ 'If
L
590
. (E. 12)
mNTlIj
= (2[ + I)N-I
L l(mjl"zjlmj)I'
(E. 13)
mj,mj giving
2
j
x 'lhwo(mj - mj) - hw J
mj) -1iw]
L
exp(tlWoMlkT)
l(m\.m2, ... • mi' .. ·1p:zjlmJ, m2 .... • mi .... )1
2
We wish to re-express this in tenns of the states la) and [b). To do so we note that J(m"m2• ... mj. ···IJ~",jlm"m2, ... ,mj, ... )1 2
ml,m2 •... ,mj •...•mN,mj
x
I
L.J, l(mjllt",jlmj)1 6(llwo(mj -
)_Imj,fflj
fill ,11I2, ... ,fIlj". /I
'"
1/
(E.7)
_
lI"nw
X (w) - (2l + I)N kT
'" 2 L, l(al"zlb)1 ,(E" -
0,'
E, - tlW)
(E. 14)
591
But this is just the result we would have had had we replaced all exponentials in (E.I) by unity. We see that we have never asserted thac lEal <: kT. In fact, we have made no approximation on this score at all. Our only real approximation is that the partition function for N spins shall be (2I + I) times that for N - I spins. A similar situation arises in numerous other places where we use the high temperature approximacion. Essentially we are saying that although the energies correspond fonnaJly to a large N, in actual fact only a small number of spins is ever really important. Restrictions on the temperature which appear because N ";:! 1023 must therefore be fictions, and we need not worry unless the energy of a small number of spins becomes comparable to kT.
F. The Effects or Changing the Precession Frequency - Using NMR to Study Rate Phenomena Hahn's observation that the bulk diffusion of nuclei in an inhomogeneous static field caused the spin echo to decay is an example of a general circumstance of great utility in applications of magnetic resonance in physics, chemistry, and biology. The existence of diffusion allows a nucleus to move from a place where its precession frequency has one value to another place where the different mag· netic field srrength produces a different precession frequency. At a later time the nucleus could move to stiU a third location, and so forth. There is therefore a frequency modulation associated with the motion. Such effects are readily treated using the Bloch equations, but to do so we lay some background by treating the more elemental frequency modulation which arises when a nucleus possesses only two precession frequencies between which it can jump. For example, consider the molecule shown in Fig. El, an N, N -dimethyl amide. The two CH3 groups have different electronic SUlTOundings giving them different resonant frequencies in the same applied laboratory static field. This effect (called the chemical shift) is discussed in Chapter 4. As explained in the figure caption, the molecule is planar, but can jump between configurations (a) and (c) if given enough thennal energy to overcome the p0tential barrier which tends to keep the molecule planar. At low temperatures the molecule is therefore effectively rigid, giving rise to two resonances (one from each CH 3 group) displaced in frequency an amount I5w. AI high temperatures, the molecule makes frequent jumps between the two planar configurations. When that motion is sufficiently rapid. the protons in either CH3 group respond only to their time average environment. the distinction between the two positions in the moleculer is lost, and only a single resonance is seen. The situation is shown in Fig. F.2 which reproduces the data of GUlowsky and Holm [F. I] for a molecule in which the molecular fragment R is CCI3· We note at -8.50 C two lines of equal intensity which gradually broaden as the 592
JI)c...........
/R N-C
JI)C/
~o
(,j
/I)c.......
80
~N_C7' If C/ ........... R )
(c)
Fig.F.la-c. The molecule N,N-dimethyl amide where R stands for either H or ClI3. The bonding structure on the left. shows an N-C single bond, but admixture of stste!! on the right gives a partial double bond, making the molecule &lIIIume a planar configuration, and giving a barrier to rotation about the N·C axis. The two CH3 groups possess slightly different chemical shifls. However, if the ReO fragment jumps over the barrie£, to achieve the configuration of (d, the Clh groups interchange chemical shifls
23"C
-1i.S"C
Jlf\A -3.S "c
IS.S "c
29.5
MAA ;J(.cr(.c 1 6
.......... 10 cps
"c
'.c
F1... F.2. The proton magnetic relIOnance spectrum of N, N-dimethyltrichloroacetamide (DMTCA) Il!I a function of temperature, at 60 Mlb measured by Gll/OWSly and /101m. The frequency scale, but not the intCfl$ity scale, is the same for all temperatures
temperature is raised. coalesce into a single broad line, then eventually become a single narrow line at high temperatures. This problem was first solved by Slichter [E2] and independently by Hahn and Maxwell [E3] using the Bloch equations. The method of Hahn and Maxwell, which we follow below, was also later rediscovered by Van Vleck [FA) and by McConnell [F.5]. Using still different means, Archer [F.6] and Anderson [F.7] also discovered the same result. Van Vleck remarks, "It is rather remarkable that although calculations of line shape based on phase interruption are of very long standing, stemming mainly from early work by Lorentz a half a century ago, the fonnula ... based on the simplest example of frequency interruption was apparently not presented until 1953 although the case is one of considerable physical interest". In order to analyze Ihe situation, we consider that a weak rotating field Hi is applied, we neglect saturation. Equation (2.87) described a single resonance 593
(F.I)
where
(F.2)
I T,
. h
0'=-+170
(F.3)
where
(F.9) (FA)
ho=Ho-w/-y
We wish to describe nuclei al two siles, a and b. Let us denote the total M~+iMy for nuclei at an a-type site as MOo and M~ + iMy for nuclei at a b-type site as Mb. MOo and Mb are then complex quantities. We define
(F.5) the displacement of the frequency from the resonance condition in H o, and introduce the splitting Ow between the resonant frequencies of the two sites. Then, were there no jumping, the nuclei at an a-site would obey one differential equation, those at a b-site would obey a second differential equation.
l Mll.( dt T2
'W) 2 ,w) 2
.2 . M,2
(F.6a)
dM. . ( Llw-- Mb+I,-H, - : - - M• +1
(F.6b)
T2
Ow =-2
or w
'w
=7 H O- 2
OMb = C2Mo.Ot
(F. 10)
oMo. = -C2Mo.Ot
(F. I I)
So we can say that as a result of jumps
where Mo is the thennal equilibrium static magnetizalion of the sum of the two sitcs_ The steady-state solution of (F.6), obtained by setting dMo./dt = dMtJdt = 0 goes through exactly as in Section 2.8, giving two Lorentzian lines of equal amplitude, one located al
L1w
When a jump occurs, the a-spins in one CH3 group leave the b-site, diminishing the b-magnetization, and go 10 an a-site, adding 10 the a-magnetization. The nature of the molecular formula for our example requires also that during the very same jump a-spins go to a b-site, but in a more general case (e.g., if we had a CDJ group and a CH3 group on the molecule and were considering only the proton resonance), the jump of a-sites to b-sites is independent of the jumps of b-sites to a-sites. By the same token a-spins will jump to b-sites adding to Mb an amount
and diminishing MOo by
dM M, - : - - + 1 Llw+- MOo+I,-H\
dt
and the amount is proportional to the small time interval ot. C 1 is a constant which depends on how often jumps occur. Of course, since MOo and Mb are complex quantities of the fonn M z + iMy , (F.8) is a vector relationship. The constant C 1 is real. The process of (F.8) will diminish lhe b-magnetization by
dMo.
&
= CI Mb - C2Mo.
dM.
dt"" = C2 M a -
(F.I2a)
and
(F. I2b)
C 1 Mb
These rates can be added into (F.6a) and (F.6b), respectively.
dMo.
Mo.. (
&=-T2
(F.7a)
+1
'W)
.
M,
L1w+2" Mo.+CI Mb- C 2Mo.+ 1,H\2
(F. 13.)
the other at
L1w = Ow 2
or
'w W=7HO+2"
(F.13b) (F.7b)
To include the molecular motion, we now assume that the molecule reorients, carrying spins from an a-site 10 a b-site, and vice versa. We assume thai when the reorientation occurs, the actual process of going over the barrier is very rapid, so rapid in fact that the protons do not change their direction. Thus if we lake a lime interval ot sufficiently long that a number of spins al b-siles jump to a-sites, they will add 10 the a-magnetization an increment oMa of the form
(F.8) expressing the faclS that Ihe eXira magnetization oMo. has Ihe orientation of Mb' 594
For our example, whenever the molecule reorients a-spins convert 10 b-spins and b-spins convert to a-spins, so that Cl = C2 = C. These equations are easy to solve in the steady state, since then they are two simultaneous linear algebraic equations for the complex quantities M a and
Mb· The result for the total complex magnetization, Ma: + iMy is
·M M M i7 H \MoT[2+r(O"Oo+O"b)!2) M z+1 y= 0.+ b= (l + O"o.TXI + O"bT) - 1
(F.14)
where 595
(F.ISa)
(F..ISb)
(F.16)
1 -=c T
The absorplion signal is given by My. the imagi.nary part. of ~F.14). The results computed by Gillowsky and Soika (F.8] are displayed In Fig. F.3. The . I bele
Fig. F.3. From the curves we see that the transition from the c~se of '~freq~e~t
jlUnping to the case o!rapidjumping occurs when raw ~ 1. Thl~ re~atlOnshlp IS widely encountend and is ojgreat importance!or numerOlls appllcatlons a/mag-
netic resonance. It describes morional narrowing. Note though the total pattern changes character when Tliw :::::: 1. there are also changes ap~nl when or6w:> 1 , _ 1 Thus in Fig. F.2 the low temperature peaks begin to broaden long
, . . h'h before they collapse and merge into a single peak. LIkewise at 19 tem.pera. lure Ihe peak conlinues to narrow with increasing temperature long after 11 has
m~w_.
become a single peak. .' . .. f In general, associated wilh motion there IS a sphtbn~ 6w, characlenstlc 0 Ihe zero-motion (long T) limit, and a physical process which ca~~s Ihe nucleus to change resonant frequency, with an associated time T descnbmg how often on the average the frequency changes. . . For example, in a solid, a typical nucleus expenences magnen~ fields i1H of order ± p/W where}J is the nuclear magnetic moment of the neIghbors, and 2f6w
, 3 4
10- 1 1 10 10'
,
b
R the distance to the nearest neighbors and the ± signs represent the fact that the field due [0 the neighbor changes sign depending on the orientation of the neighbor. Consequently the resonance of the nucleus in question is spread above and below the value w:::: "(Ho. In a solid, a given nucleus has many neighbors, some aiding, some opposing the external field. Thus instead of having twO peaks as in the case we analyzed, we get a smear with a width somewhat broader than would result from having only one neighbor. We thus make the rough identification 6w Z,,(/J
2:::: R3
(F. 17)
where Z is a small numerical factor representing the fact Ihat there is more than a single neighbor. The process of self-diffusion enables the neighbor atoms to move, so that one neighbor is replaced by another. In the process, the resonance frequency of the nucleus undcr observation may change depending upon whether a neighbor is replaced by a nucleus whose moment is oriented in the same direction or in the opposite direction from the magnetic moment it replaced. If T m is the mean lime a neighbor sits before jumping, we surmise T:::: T m
(F. 18)
At low temperatures, then, we expect a nuclear resonance line width order 6w/2 given by (F.17). As we increase the temperature, we expect T m to become shorter. When (F. 19)
we expect the nuclear resonance line width to begin to narrow. Exactly such effects are seen, as is discussed in Chapter 5. Chemical exchange can be studied by its effect on the nuclear resonance absorption lines. For example, the spin-spin coupling in a molecule gives rise to structure in liquids, as illustrated in Fig. 4. I I. In liquid CH3CH20H, the OH protOn splits the CH2 proton resonance structure if the OH proton remains on the molecule. But if there is rapid chemical exchange of the OH proton with other protons in the liquid, the structure may be washed oul [F.9].
G. Diffusion in an Inhomogeneous Magnetic Field
1
2
o Ow " I h"n' I' h resolution .'ig. F J The avcnging by molecular reoricn~ll~ion of Achcmlca S l In a ug" ( d NMR ~pec~rum" It is assumed lhlll mu~ual reorientAlion oecurs belw~n ~wo Sll~ ~ a~.... b) which arC equally populaled and which hAve resonancc frequencIes separat y radians in lhe absenec of thc exchangc
596
In Section 2.9 we explained how diffusion in an inhomogeneous static magnetic field causes a spin echo to decay as the time between pulses is increased. The cause of the decay is the possibility for a nucleus to change its precession frequency by diffusing to a different point in the sample at which the static field is different owing to the inhomogeneity in the magnetic field. Such a situation is closely related to that treated in Appendix F in which the nucleus had two
597
possible natural precession frequencies. For the case of diffusion in a magnetic field, there is a continuum of magnetic fields. For simplicity, we assume the magnetic field, though inhomogeneous, has axial symmetry so that
8H
H(x,y,z,)=Ho+z az
Mz. and My arising from precession. If we transform to the reference frame rocating at the precession frequency, we eliminate the precession and can write in this frame
8~::
(G.l)
In Appendix F we described the process whereby spins switch between sites of different precession frequency by the nile term (F.12a) and (F.l2b). When diffusion can take place, the possible precession frequencies form a continuum as expressed by (G. I). The usual way to describe diffusion is by means of a diffusion equation. The use of a diffusion equation in conjunction with the Bloch equations was introduced by Torrey [G. I ]. We shall follow a rreatment which is a slight simplification of his. Suppose we had a homogeneous static field in the z-direction, had no applied alternating field, and by some means had produced a nonuniform M z , as in Fig. G.Ia. Let us suppose T t is infinite. TIlen the total z-component of magnetization cannot change, but as a result of diffusion the region of magnetization will spread as in Fig. G.Ib, eventually leading to a uniform M z throughout the sample (Fig. G.lc). (Recall that the symbol M z denotes the magnetization density.) The process of Fig. G.I is described by the equation 8Mz = DV 2 M (0.2) at ' where D is the diffusion constant If the magnetic field were uniform, and there were initially also :t- and ycomponents to the magnetization density, there would also be diffusional effects for Mz. and My. But since the static field causes a precession, two effects are present: (1) changes in Mz. and My arising from diffusion, and (2) changes in
'" M DV2
(0.3,)
z
a;:y = DV2My or
(O.3b)
8M+
--=DV 2M+ at
(0.4)
using the definition M+ == Mz. + iMtI • . Since M+ describes a two-dimensional effect, (GA) treats vector effects. It IS analogous to (F. 12), if we bear in mind that M+ is a function of position. We now proceed as in Appendix F to recognize that when the static field is inhomogeneous, we must now include the effect of the spread in precession frequency. We therefore add the precession driving terms giving
aM+(~ y, %, t)
= _ i-yh(x, y, z)M+(x, y, z, t) _ M+(:t. y, %, t)
T, (0.5)
where
h(:t, y, z) = H(:::. y, z) - H o
(0.6)
and Ho is the spatial average field over the sample. Substituting (G.I) for h(:t, y, z~, w~ ~btain an equation describing free precession in a static magnetic field which IS mhomogeneous but possesses axial symmetry:
8M+
(8H) M+ - -M+ + DV 2M+ T2
(0.7)
- - '" -i-yz i)t 8z (.)
M,
M,
M,
o
L
(,)
(b)
o
L
o
L
Fig. C.lao<. A sample in II static field in the z-direction is lnllgnetized in the z-direction. The magnelization density M~ is IIllSUmed initially to be nonuniform. Figure G.ta shows the initial M~ as a function of the :a:-coordinate, the sampJ.e extendins from: = 0 to z. = L. As a result of diffusion, the magnetization density at a later time (Fig.G.tb) has spread throughout the sample, with a diminished peak. Eventually, the M. becomes unifOnlJly spread as in Fig. G.le. Note that if T 1 is infinite, the lotal z-magnetiution, which ill the area under the curve of 11-(.(:), is the same in all three casn
598
Equation (G.7) incI~des the natural T2 effects as well as the T, effect. [The fact that T} effects are Included can be verified by examining (2.86) with HI set equal ~ ~ro). Note that since aH/8z is evaluated at the origin (:t = 0, y = 0, z = 0), 11 IS a constant The only explicit dependence of (G.7) on position in the sample is therefore in the first term on the right-hand side. If there were no diffusion (D = 0), (G.7) could be solved simply since over a layer of constant z the equation describes precession in the rotating frame in a constant field h, together with decay at the rate T2. The result is
M+(r,t) = M(r,O)e-I/T'e- i-r z(81l/8z)t
(0.8)
=
where M(r,O) is the complex magnetization density at t O. Suppose that M(r,O) were uniform, as would be the case if it were prepared by applying a 7r(l pulse to a sample initially magnetized to thermal equilibrium along the 599
BH
static field. (We here neglect the slight variation in thennal equilibrium static magnetization produced by the small inhomogeneity in a sialic field). Consider
-V-T
Bz
th, z, and z + L},z where az is a small distance. Since these three planes are equally spaced. their precession frequencies are equally spaced. Swting in phase at t = 0, the magnetization at z+ L1z develops a lead in angle over that at z, that at z -.dz develops an angular lag relative to the the magnetization density al
Z-
Thus at
1I1+(r,t) = Moe-tjT2e-i.,(8ff/8:)IA(t)
A dt
A
'=
A(O)exp
.
We now use (G.12) to describe the development of M+ in lime following T:
M+(r,t - T) = M+(r,,)exp ( _ (t;2 T»)ex p [
( . BH
x exp -I"(z Dz (t - T)
M+(r,2T) = -Moexp(-
or
(G. 10)
BH)"'] 3" [-v ("(a;
(0.11)
M+(r,t) = Mo exp (-tIT2)
xe.P[-D(' ~~)":lexP[-i'Z( ~~)t]
Mo exp ( _ ;2)ex p (
x ex p ( -i"(z
~~ T)
7
(G. 14)
t = 2T), the complex phase
~)exp[ -v( ~~J2;3]
(G.15)
This is Hahn's famous result. his imponant to note that the diffusion tenn after two intervals of T has an
e~pon.ent of ("(DH/Dz)2V(2,3f3) rather than ("(DHl8z)2V«2T)3/3). That is, the ~Jffuslve ~hase loss takes place independently in each time interval T. This fact IS the basIS of the technique of Carr and Purcell fG.2] discussed in Cha t 8
who at h · f \. , p er , net at lone app les a sequence of 'II'" pulses spaced 2, apan, one obtains a sequence of ~~s, and by making the pulse spacing sufficiently close, one can make the diffUSive Joss of magnetization as smaU as one wishes relative to the T2 tenn.
(0.12)
for the magnetization following an initial 7fn. pulse. This equation describes the development of magnetization with time from its initial value Mo at t = O. We now need to consider what happens if we apply a 7f pulse at time T producing a rotation about the y-axis. The magnetization density just prior to the pulse is '=
-n("( ~~Y(t ~ ,)3]
)
Substituting (G.~3! we see that at t - , = T (or factors cancel, glvmg
(G.9)
The constant A(O) we incorporate into Mo giving
M+(r,T-)
Bz
t = T+
x exp (iJr) exp (i"(z ~~ , )
When this solution is substituted in (G.7) we get a differential equation for A(t)
t2 _v(/H)2 Dz
BH
+ "(Z-T
M+(r,T +)=Moexp ( -T2T) exp (-V (BH)' T') 7 Dz "3
change for the resultant magnetization at z. Therefore, we expect that diffusion will not affect the phase of the development of the magnetization at z, but will affect the magnitude. As long as we are not near a boundary, every plane z has planes symmetrically Llz above and below with phase lead and lag, respectively, from which magnetization diffuses to z, the phase lead or lag depending on .az independent of z. Thus we expect that the diffusion-induced decay will be independent of z. We therefore try a solution
'=
'II'"
I
magnetization al z. the lead angle and the lag angle growing progressively with time, bUI always remaining equal. When spins from both z + L1z and z - L1z diffuse to z, they will do so in equal amounts. Thus they will add in equal amounts increments which lead as increments which lag, giving no net phase
2. dA
into
-v("( ~~Y~3) (G.13)
The 11" pulse leaves My unchanged, and changes M z into -Mz . This is equivalent to changing
H. The Equivalence of Three Quantum Mechanics Problems In this appendix, we show the equivalence of three quantum mechanics problems:
1.
A. pair of identical spin ~ nuclei coupled to a strong static field, interacting wilh each other by a coupling of lhe fonn l:S:, where z is the direction of Ho.
2.
The same p~le~ as (I) except the spin·spin coupling is dipolar. A nucleus Wlth spm F = I, coupled to a strong static magnetic field, with a quadrupole coupling which is axially symmetric about the direction of
3.
110. Defining
WQ
= "(Ho the three Hamiltonians are then
600
601
i
(2)
'Hb = -rIWOUl
+ Sl) + M(3I z S. - I· S)
(H. I)
Table H.I
(H.2)
SLale
•• .p
where we have kept just the secular part of the dipolar coupling 2 (3) 'Ho = -hwoFz + rlc(3F; - F )
(H.3)
p. pp
Table B.2
Table H.3
Slate
Slate
Energy
f" f" 1/>\,-1 f ..
-,
Energy '-"'0
-wo + a/4 f" -a/4 f" WO + a/4 lPl,-1 -a/4 f ..
+ a/4
-a/4 -0./4 WO
Energy
+ a/4
"'0
+ bj2
WO
+ b/2
0
For problems (1) and (2) we consider two basis sets (I)
lmJffiS}
. We use a notatIon
where
rnJ =
±!
mS =
±!
(HA)
• to wnle . wave f ' for m/ = +'2I mS = unctIOns
lrIJ
~iI
(2) The singlet and triplet states Defining F = 1+ S we have the commuting operators IFl tal angular momentum quantum numbers F associated with eigenvalue of F z giving four states
F =1
M = 1, 0, -1
F=O
M=O
2
and F z with toand M, the
IFl 2 ,
(the triplet states)
.
(H.S)
combination of them is also an eigenstate. In particular, the two un normalized states u{J + {JOI and Ol{J - {Ja are also eigenstates. But when nonnalized these are nothing but the two states 1/;10 and ""00 of (H.6) Therefore, we call al;o rake as the states those shown in Table R2. We turn now to the second Hamihonian
Ji b ::= -nwo(Iz + Sz) + hb(3Iz 5 z - [. S) First. we nOle that the tenn I· 5 can be rewrillen utilizing the faci Ihat IFj2::= (I+ 5)2::= 1 2 + S2 +2[· 5 (RIO)
'H, = -f~o(l, + S,) + M[3I,S, - ~(IFI'
In tenus of the states Im/ffiS) we have
VJIl
=
0'0'
1
>1'10
= .j2(aP + pal
~I,-I =
(H.6)
pp .
We note that the states l/JlA-f are unchanged if we interchange the labeling of the two spins, whereas tPoo changes sign. They are thus eigenstates of the pennutation operator with different eigenvalues (+1 and -I respectively). On the other hand, an operator such as Iz+S z is unchanged by the pennutation (Sz+Iz --+ Iz+S z ), hence commutes with it. Therefore, I z + Sz has nonzero matrix elements only between states with the same eigenvalue of the pennutation operalOr. Thus, its matrix elements between the singlet and any of the triplet states vanish. To solve the first Hamiltonian, H a , we note immediately that both I z and Sz commute, hence we can take as exact eigenstales the individual states lmlms)· The energy levels are then immediately (H.7)
It will be convenient for all three cases to express the energy eigenvalues in units
of frequency. Wi. Wi
So
== Eil h .
Will/illS
::= -wo(ml + ms) + amlmS
(R8)
- III' -151')1
(H.ll)
Now•.up through the te~ I z 5 z this Hamiltonian looks just like H a with 3b replaCing a. We could take either 1/;11111115 or 1/;F,M as eigenstates for this much of the Hamil~onian. For the last tenn, since III 2 and j51 2 are the same for all st.ates, they SImply add constants to the energy. The tenn 1F1 2 is the significant difference from the Hamiltonian H a . However, if we take the states ""F M. they are also eigenstates of 1F1 2. ' Utilizing that
(FMIIFI'IFM) =F(F+ I) (FMIIII'IFM) = ~ = (FMIIS'IIFM)
(H.12)
~(IF'I-II'I-Is'll = ~F(F+ I) - ~
(H.13)
0'
we get Table H.3. At this point it is useful to note that one can fonn a simple check on one's calculation of the individual eigenvalues from the observation readily proven from the explicit fonns of the Hamiltonians, that ' T1' {'H.} = 0
,
T, {'H,) = 0
,
T, {'Hol = 0
.
(H.l4)
Thus, adding the diagonal elements must give zero. The energy columns of Tables RI-3 all add to zero. The last Hamiltonian is
(H.9)
Ji c ::= -hwoF: + hc(3F; -
1F1 2 )
See Table RI. We note that the states Cli{J and {JOI are degenerate. Thus any linear 602
603
for F = 1. Obviously I Fl 2 and F; commute with il so we immediately take the stales .pFM as eigenstates. getting Table H.4. .. . For all three Hamillonians. we could induce transitions by applymg a transverse alternating field. (H.15)
Hz(t) = Hzo cos wt
=
=
by (H. 17)
, .6w=6c
l;reqllcneics of tra.nsitions
Table 1f.5.
Stllte
lIa.miltonill.1\
Energy
Fl'e<:lucney of the to 110) transition
III) 1l.
WO
11, 11.
WO - 3b/2 WO - 3c
a/2
Frequcney of the
110)
to II - 1) tf1lnsition
wo + 0./2 WO + 3b/2 ... +3<
Suppose we take the Hamiltonians 1-£0.' 1-£b. or 1-£1:: as an unpe~urbed Hamiltonian 1-£0, with eigenstales U/ and energies E!. We can then co.\S1der .the effect of any combination of pulses by adding their lIme-dependent perturbation Hp(t) to give (H.IS) ?itt) :?io + ?ip(t) Then. the wave function 1J1(t) can be written as iw L = l:>/(t)u/e- /
wet)
(H. 19)
I
Since the coefficients a/ are independent of time if H p = 0, the foml (H.~9) is the interaction representation. Substitution of t/J(t) in SchrOdinger's equauon leads in a straightforward way [see (5.78,79)} to
da" dt
604
=.!. 2::a/(l'I1-£p l1)e-iW,t 1l /
i
.
datO
dt =
- Wto)t]
i
.
- 'h{a ll (lOI1-£ppl)exp[l(w lO -wlllt)
+ al _I (lOI'HpP - I)exp [i(wlo - WI dal_l
i
_I )t]
.
~ = -'halo(1 - tI'HpllO)exp[ -1(WI_I-wlOlt]
These equations are identical for all three cases provided one notes the relation (H.17) between the three coupling constants; a is equivalenl to 3b or to 6c
for lhe three cases. Table n.4
(H.2t)
(H.16)
where Fz = [z + Sz for cases (a) and (b). We have already shown that Fz has no matrix elements joining the F I and F 0 states of cases (I) and (2). Thus, the only lransitions induced are between Ill) and [10) and between [10) and'p -I). We can use the tables of energy levels 10 calculate these two freq~encles. See Table H.5. These frequencies correspond to a pair of lines separated In frequency
L\w=3b
.
The other aFAI's obey the equations
da
1-£1(t) = -"(llH zo F z cos wt
,
aoo = constant
ll dt = -'halO(l l[1-£pjlO)exp [l(Wll
producing a time-dependent perturbalion, 1-£1(t), given by
.6w=a
Since (l'J1-£pjl) vanishes between F = 1 and F = 0 stales, for cases (I) and (2) of coupled pairs of spins we get. using the notation aFAI
.
Therefore there are no stalic or dynamic observations one could make which would enable one to tell which of the Ihree systems one had. One may ask. what if one had chosen to use inSlead the Im/ms) representation? Clearly the physical result should not change. All that will change is thai one now has four instead of three "allowed" transitions (all' to ap or po'. and fJP to afJ or fJa). However. if one were to do the problem explicitly one would need to be mindful of the exact degeneracy of the afJ and fJa stales for case (I), which guarantees coherent effects. If one uses the product operator formalism all these things are automatically taken care of since we do not specify a representation when we use operator algebra. One could. of course, start with the representation lm/ms). formally express the states in terms of the states IFM), solve the problem using the IFNI) states. then transform back [0 jm/ms). An impor1ant practical consequence of the equivalence we have proved is that if an X(?r(2) ... T ••• Y(?r(2) pulse sequence refocuses the coupling aI~S~, it will also refocus the true dipolar coupling. As is shown by (10.59), the effect of a general quadrupole coupling to first order is always of lhe fonn of ?le. Thus, for a spin I nucleus, 'He describes the first·order quadrupole effects. By analogy to the echoes described above, we can refocus the first-order quadrupole coupling of a spin I nucleus using lhe same X(1fn.) .. . T . .• Y(1J'fl) pulse sequence.
I. Powder Patterns
(H.20)
In Chapter 4 we saw thai, in general. Knighl shifts are anisotropic. The same thing is true for g-shifts encountered in electron spin resonance and discussed in 605
Chapte r 11. These features give rise to a resonan ce frequency in (4.187), goes as
W
which, as shown
0.1) to the cryswhere (8, q,) specify the orientation of the static field Ho with respect giving rise ly, random present are (8,,p) tal axes. For powder sample s, all angles singularieristic charact quite r, howeve to a spread in resonant frequencies with, powder called are spectra Such . ties from which wo • wb. and We can be deduced We [1.1]' d Rowlan and ergen B/oemb pattern s. This problem was first treated by we so, doing In s. powder tum here to calcula ting the intensity pallems for such some or crystals liquid are omillin g discuss ion of fascinating systems such as crystals and form of oriented polyme rs which are concep tually in between single random powder s. that the We express the random orientation of the crystallites by saying by given is probabi lity dP that Ho lies in any infinitesimal solid angle dfJ (I.2)
dP "" dfJ "" sin 8 d8 d.#47r = d(- cos 8)d#41f
4,
Introdu cing the variable z defined as (1.3.)
z=-oo s8 we have that dP is
dP = d~dz
(l.3b)
4~
The advanta ge of the variable s (q" z) is that equal areas in the spond to equal probabilities. Inttodu cing fJ defined as fJ
==
W -
rlrz
plane corre--
(1.4)
Wo
we have
n "" W(I sin 2 8 cos2 ,p +wb sin2 (J sin 2 q, + We cos2 8 2 = W a + (wb - w a ) sin
,p + [(we
2 - w a ) - (Wb - w a ) sin q,]z2
(1.5)
frequency, fJ. This express ion can be thought of as detenni ning lines of constan t have we in the rlrz plane. Solving for z2 2 -wa - (Wb -wa )sin q, 2 ~~ z= 2 (We - wa) - (wb - W(I) sin 4J q, at constant SO that it is straight forward to calcula te the curves of z versus er, then, frequency in the ,p-z plane. We illustrate such curves with Fig. 1.1. Consid lity of probabi the (1.3), of result a As dfJ. + fJ nand two curves at frequencies area the just is range cy frequen this within cy frequen finding a nucleus whose the also is lity probabi the But curves. two L1A in the rlrz plane between these have we Thus J. 1(fJ)df nonnali zed NMR line intensity
n
606
06
0.'
0.2 l'lSlJ.
-2:,j L~.,-o~o'---~20I-'--,,,:o--,,~-,:;--f-;;:;;--:-::--:~. 160 le.O 80
90 100
120
'"
¢lldrgrr ul pJ F'ig.l.I. Curves of constan t frequene Y .In the..t ane for" general anillOtropic chemic,,' '1'": or Knight shift
l(fI)Ll fi = LlA
~
~
. This relation ship shows that if we divide the total frequenc y range Into equal be frequency interval th s. e area tween successive curves of constan t fre uenc . . P~portlofnalto (he in~ensity at that frequency range, giving us a simple ~Phrc~: picture 0 where the Intensity is high, where it is low. We case now describ e how to calcula te len). We first treat the simplest . I ' case. general the do then try, symme axla the same Axial symme try means that two of the principal frequencies are . . easIest ' IS (e.g,w(J =w,,~w orw =w t The ca Icu Iatlon et) if we pick Wb, c.. C (I c' . th 8 _ 0 aXIs to be the symme try axis (W(I = Wb) since then axial sYrnme try e .. . h gIving means that the q,..dependent teons of (1 .5) vams
n=
W(J
+ (we -
w(I)z2
0.8)
q,-z
plane are straight Jines at Therefo re, cu.rves of constan t frequency in the ~n~~ti: t~~~' 1.2). The area, L1A', between two curves differing in frequency
LlA' = 211"Llz
(1.9) vertical the whe~ the 21f ~mes from the total range of ¢J, and where Liz is spaclOg approp nate for the range Lin :
Llz =
(afla,) -
Ll
fl.
Such a strip occurs both for
(1.10)
O<8
1 .0 § §
O. 0.8
0.8
0.7
0.6
FIg. L2. Curves of COfl5tan~ r~uenc y in thf: t/J-z plane for axially !Jymmct-
ric chemical or Knight shifts. " is the nonnl\li~f:d frequenc y
defined by (1.16)
~5 y
z
O.•
0.4
l--- ---- ---- ---i 0.4
w
G.OS
u
~
~
W I(y)
(LIt)
.cl.A = 471'".cl.z
I
an "" 2J(n
wa)(w c
(Ll2)
wa)
giving
Llfl XP ==C"") ,(nflF5 I(fI)Ll fl = LlA/4. = Ll, = '2J W(JW eW a v
(Ll3)
n
n
between We and We_ The intensity is zero for outside This Cannula holds for rity (infinite this range. This famous line shape has a mild (square root) singula = W o ' This frequency corresponds 10 Z =- I, or but integrable intensity) for The infinity () = 71'"/2. when Ho is perpend icular to the c (i.e. symmetry) axis. ism. mechan ing broaden al disappears if there is an addition we> W a Note that this formula holds true whether We >w(l or We < Woo If with end, high the on We to the line extends from Wa on the low end
n
$ O$we
We
We
< Wa
$ O:S Wa
(Ll4)
~o~~i:;~t
=-..!2~
if we comput e the total area of the line by integrating I(y), the result
I
Jo
~~;.~~~.g that
(Ll9)
I(y) of (l.I8) is properly normalized. This line shape is shown in
• ..l. To deal with the more general case that w..l. a .,. Wb .,. We, we use two dlmen . b s· I Ion ess vana les y and f. We first rewrite (l.5) using (1.16) and (I.17): {)-w a y""
We-w a
n-
The W a is of the total frequency interval W _ W for the fraction that (I. e • as second variable j is defined j""Wb -Wo
We
Q.20a)
- Wa
Ilyl
(the symmetry axis is at the high frequency end of the (1.15)
argument of In this case. both (0 - w o) and (we - w o) are negative. so thai the the square rool in (1.12) is still positive. define It is useful to introduce a dimensionless, variable y in place of n. We yas 608
> W a or
~t~
I(y)dy = I
From (1.8) one readily finds
8,
We:
(Ll7)
2~
frequency function of cos 8 (Le. z). Thus, the tOlal area LiA which has a given range is twice l1A':
If. however. spectrum).
see that y is always positive whether
a to I:
n ""
4' ldegnul
Wo
From. (I.13) and (1.14) we We <Wa • and extends from
I I(y)Lly "" -Lly or
O'--~~~~~~~e_J
o
(Ll6)
W a , so y is a dimens ionless f ~e val~e. y "" a corresponds to requency e that 0 is from ,h distanc the of fraction coordinate glvmg the W a end of e . . . Th I the frequenc . t en, diViding numerator and denomi nator of (I . 12) by y In erva. we-w awege t
f--- ---- ---- -10 .1 --- --- --- -
O-Wo we:-wo
O$y$ I
1-- ---- ---- --1 0.' 1--- ---- ---- -10 .7
0.2 f---
y=
o
0.2
0.4
Y
0.6
0.8
1.0
Fi~ I.J. NMR line shape ror An aXially symmet ric chemica l or ~night shift tensor. NMn. inten_ sity 1(1/) versus norm"li~cd frequency II
609
,., 1.0
We adopt the convention that WI! lies between w'" and We.. but allow either Wo > We. or Wo < We.' Thus W/I can be either the low or the high end of the frequency spectnlm. With these conventions we have (1.201»
OSfSI
Note that f tells what fraction the total frequency interval We. -W/I lies between Wo and Wb' In other words, Wb divides the interval into twO fractions f. between w'" and Wb' and 1 - f. between wI! and We.' Then we gct from (1.5)
y=
f
sin 2 ¢ + (I -
f
sin 2 ¢)z2
(I.2l)
":1g.l4. Lines of e.onlllant frequene.y tI in one pottlon of the >-z plane for 1=0.75
, 0.'
0.'
0.' 02
0.4
yzf~O.lS
0.2
and from (1.6) [or by rearranging (1.21)}
f sin 2 ¢
(1.22) 1- fsin 2 ¢ From equation (1.22) we can draw some simple conclusions about the nature of the curves z(¢) at fixed frequency y. Since z2 = Y -
sin2 ¢=sin2 (¢+11") the curves for
11"
(1.23)
('II" _
¢)
(1.24)
the curves for 'II"{l ~ ¢ ~ 'II" are mirror images about ¢ = 'II"{l of Ihose for 0 ~ ¢ ~ '11"/2. Therefore, we need only consider the curves for 0 ~ ¢ ~ K/2. Since (1.22) involves z2. Ihe curves are an even function of z. Thus, we need consider only posilive values of z. The net result is that if we calculate any areas for the octant 0 ~ z ~ I. 0 ~ ¢ ~ 'll"n. we must mulliply the net result by a factor of eight to get the full area. To get the axially symmetric case with Wa = WI!. we simply set f = 0 which gives then z2 = Y
4' ldtl}/"usl fl=w a
For the frequency y = 1 (fl = we.)
~ ¢ ~ 211" simply repeat the curves for 0 ~ ¢ ~ 11". Moreover.
since sin ¢ = sin
O~OC--20;,---l--::",~-~. :. -l--~..~J90i...
.
(1.25)
z=1
(1.28)
ind~pe~dent of ¢. Therefore ¢ = W/I is the origin, ¢ = wI! (y = f) is the line which Intersects the z = 0 axis at ¢ = 'll"n, and W = We. is the line z = I We h v . a e marked these curves on Fig. 1.4. To calculate I(y) we now consider two lines in the first octant of the titplane, to frequency n, the other to fl+ L1n. Then, recallin; I(y) L1y IS gIVen by eight times the area, L1A', between the two lines at positive
on~ c~rrespond~ng
z. I(y)Lly = 8LlA'
(1.29)
Foc:using on two such curves, we examine Fig. 1.5. We look first at the small regIon bounded by lines at y = constant and y + £1y = constant. The area £12 A' of the small shaded region. essentially a parallelogram, is £1 2A'-Ll - z Ll." . (1.30)
Figure 1.4 shows the lines of fixed frequency y for the first octant (0 ~ q, ~ 7rn, o~ z ::; I) for the case f = 0.75. For some values of frequency y the z(¢) curves go from ¢ = 0 to 7rn. Others intersect the z = 0 axis at smaller values of q,. The curve which divides these two classes has a y which has z = 0 when ¢ = 7r/2. From (1.22) that is (1.26.) y=f .
Th . Ll"· . e spacmg Y'IS am·mary. but the spacing £1z must correspond to the frequency Interval. hence is given by (1.10). Therefore
which means
From (1.22) we have (1.26b)
{l=WI!
or
0%
oy
Equation (1.22) also shows us that at the origin (¢ = 0, z = 0) y =0
Ll 2A' = 0% L1y Ll¢ 8y
(1.27)
I
I
= 2z I -
1
f
sin 2 11 =
(1.31)
I
I
'2 ~.;7;=1~f"'S'Cin"j'=;¢:-';7Y=-=;f"s"'in"'=¢ I
=2.jY :-.;r.I===f~s~in'j"=;¢~.;F.l=;(f'7/"y):=""'·nr.,¢
0.32) 611
6'0
or
FIg. LS. Curves constant frequency in lhe ~, plane for two nearby frequences nand + .dO
,
n
Since y
p'
s;: 1, f < fly,
=I
(1.39)
.' = Ily
giving I( ) _ 4 Y - y/(1
=
and we can identify
·1'
Jo dX/
f)
-::-YVm(14~/)
'1'
Jo dx /
(fly) - I . , 1 -l_/Stnx
1
l(l - y) . , - (I _ f)y sm x
(1.40.)
(1.40b)
• Therefore from (1.29-31) we get
8
I(y)m-
'mJ..
d.
(1.33)
2.,fY 0 ../1 - Isin' ."/1 - (fly) sin' •
sin 2
In this equation, tPmax is the maximum value of ¢ for the particular frequency, y. Thus, as can be seen from Fig. 1.4, (1.34)
If y
< I, tPmax is easily found by sin2
7.
(1.35)
7rn.
a
,
O<::Z:~1f(l
dx
J where
J
=
F(n, k)
a
(1.36)
1
o Jt_p2sin2::z:Jt_q2sin2::z: =
.,fY
1
"/1-sin'."/I-(ylf)sin'.
J0 ../1 - ysin' .. dO' ../1 - (y/f)sin' ..
'1'
1 = ,jJ
= 7rf2. it is also called complete.
diP
(fly)
(1.38)
(1.42)
Apart from the change in the inlegralion limit. Ihe right-hand side differs from the left solely in imerchanging y and f. AI all places, therefore. for y < I (n < w6). we can transfonn (I.40b) to get
4 ;cr::w
J
d.'/
y(I-f).,.,
1
- (I - y)f Sin
'I'
o
'1'
0'
J-r.'==;=e=r.'7.-~~F. .,fY 0 ../1 - sin' ."/1 sin'. I
drjJ
·/2 (1.37)
2 2 o Jl - k sin 0'
(1.41)
.
~' is chosen so that when ~ = O. ~' "" 0, but when ~ "" ~max, ~' "" rrl2. With Ihis substitulion, it is slraightforward to show that
I(y) =
4 ./2
612
(~)
~F O'IV~
d.
is an elliptic integral of the first kind. If Thus if y > f en> wb) we have
I(y) = -
~' "" Ly sin 2 ~
I rbm.x
setting z = 0 (1.22):
If ¢mu = the integral turns out to be a complete elliplic integral of the first kind. GradshU!yn and Ryzhik [1.2] show thai if O
For the case of y < I (n < W6), where the maximum value of ~, ~max, is less than 7f12. we can transfonn Ihe inlegnl by a substitution of a new variable ~' defined by
m Jo d.'/
1 _ (w - wo)(we - wb) sin2 (we il)(Wb wo)
~'
(1.43)
An altemalive nOlalion commonly encounlered for Ihe complete elliptic integnl of the firsl kind is K(m) defined as 613
tr/2
K(m)'
Jo Vl-msin I
dx
2
(1.44)
x'
Using it we can write for y < f (12 <wII)
_
ley) -
and for y
4
,Jf(l
>f
y)
K(Y(l-f)) (I
(145)
y)!
.
0
(n >W/I)
K(I-y)!) j)
_ 4 ley) - lY(1
which brings out clearly the relationship between the frequency regions on the two sides of Note that y is the distance in units of normalized frequency from the "lower" edge (y 0), whereas 1 - y is the distance from the upper edge. A similar relation holds for f and 1 - f. The argumenls m of the two ponions of the spectrum both approach m = I al = Wb' As m goes from 0 [0 1, K(rn) goes from about 1.57079 to 00, blowing up at m = L Thus, the line has a singularity at = Wb, but is finite everywhere else. The line shape is shown in Fig. 1.6. Note from (1.45) at y = 0 and y = I the intensity is
=
n
n
[(0)·
~K(O)
[(I).
~K(O)
(1.46)
so that [(I) [(0) •
If
VI::!
(1.470)
[(We) = JWb - Wa [(w o ) We -
w"
0.47b)
It is clear that from such a powder pallem one can very directly detennine wo. Wb. W c since they occur at the edges and at the peak.
2
2
Ii' £Jfn.
=_ hH + e qQ (3cos 8-1)(3m2 _1(l+I)) .., om 41(21 _ I) 2
with transition (I 45b) .
y(1 - f)
w".
Another importanl class of line-broadening mechanisms is that involving electric quadrupole coupling, a subject first treated by Fe/d and Lamb fl.3). As we saw in Sect. to.5, if the electric field gradient is axial, then in a strong field the energy levels obey (10.63)
f~uencies
wrn,rn-t
wrn,m_1 = (Ern - Em_i'Jlh
+
= WO
2
e qQ 4[(21
2 3cos 8 - 1 3(2m + I) I)h 2
(1.48)
This leads to the angular pattem for anyone transition which is similar to that of Fig. 1.3, with a singularity at 8 = rrn.. However, for each transition (m. Tn - 1) there is a cOlTesrnding one between levels with the m values reversed [e.g. in addition to ('2',~) there is (-~,-i)], This occurs at (m' + I,m'), where m l = -(m + 1), leading to replacing the factor (2m + 1) by (2m l
+ 1) = -2(m + I) + 1 = -(2m + 1)
(1.49)
This reverses the sign of the coefficient of (3 cos 2 (J - 1) in (1.48), so that the pauem is reversed about w = we (Fig.I.7). Bloembergen [1.4] noted that if 1 is ~. ~. elC., the term (2m + I) vanishes when m = so that the transition has zero first-order quadrupole splitting. He shows that it is shifted in second order by an amount
-!.
hLlw
(i, -i)
=.2..
4
(21 + 3) e Q2 q2 (I _ 9 cos 2 8)(1 _ cos 2 8) 64 4['(21_ I) hwo = A(I - 9 cos 2 OXl - cos 2 (J)
(1.50)
1.0 0.'
0.' Ily)
0'
0.'"----
_ e qQ.
o~o--'O:':.':--O;';.':---;;O.•,--JL.,o.'::,--;1.'=-O-
,
614
Ag. L6. NMR intcn3ity l(y) versus normaJized frequency y, for I = 0.70
0 e qQ. "'-"'0 21'1 '1'1 Fig. L7. Powder line shllpe for 1 i1 NMR b<-oadened by II first-order, "xilllly symmetric c1edric field gradient, The singula~ities at .... - wo = ±e'qQ/4h have Ixoen made linil.e by convolution with a na~row ~adening funo:tion,. This ~a.r: cur~e "Is<;> "rises in the powder line sh"pe of " ""Ir of dlpolar-c:oupled Idenlical spm 2 nucleI, see (1.51)
=
615
t
FIg. U. The line shape of the ~ to - NM R transition resultinr; from the powder average of an axially symmetric eledric field r;radient. The broadeninr; ariseS from treatins the field yadient to lIeCOnd order. A. is defined by (1.50)
I(ul
produces an infinitesimal change in ¢ - hence as the infinitesimal time goes to zero, so does the change in ,p). Thus the wave function just after the sudden change is identical to its value just before. So if we divide time into intervals t .. t2' ; .. ,tN with time-dependent Hamiltonians in each interval 1i lt 1i2. etc., we immediately get that at lhe end of the nth interval: 1/l(t) '" exp (
-*1ill l n )
... exp (
-*1i2l2 )ex p ( -{1illl )¢(O)
(1.3)
If the 1ik 's commuted, we could write this answer as w
t/J(t) '" ex p (
If one expresses this relationship in tenns of the variable z of 0.3a), and utilizes (Db), 0.7), (1.10), and (1.1 t), one obtains the angular pattern of Fig.U. The case of lhe second-order quadrupole broadening of the transition has been treated by various aulhors, and is summarized by Gerstein and Dybowski [1.51· The first NMR paper to analyze a powder paltem line shape was Poke's treatment (1.6) of a pair of spin! nuclei coupled by the dipolar magnetic fields. This pattern is the same in fonn as that of the first-order quadrupole splitting for a spin ~ nucleus except that the singularities occur at
(!' -!)
3 'Y2h
Iw-wol "'4-;:J
-* ~1iktk
,,(t) = exp ( -
~/
(1.4)
)t,b(O) or as
(1.5)
h(t')dt') ,,(0)
However, in general
[hi, hj]
,,0 ,
(1.6)
so one is left with (J.3) as the best one can do. It is useful to have a more compact way of writing this result. This can be done by a method due to Dyson {J.I]. He introduced the concept of a time ordering operator, which we will denote by To (D for Dyson, T for time). Using it
(1.51)
TO exp ( -*1i1 tl )ex p ( -*1i2t2) '" exp ( -*1i2t2 )ex p ( -{1i1 tI) (1.7)
0'
J, Time·Dependent Hamiltonians
Toex p (
We have seen that when the Hamiltonian is independent of time, we can fonnally solve the Schrtldinger equation (1.1)
by use of exponential operators
,,(t) = "p ( -
*
ht) ,,(0)
"'ex p ( -{1i3t3)exp(-*1i2t2)exp(-*1itt1)
(J.8)
In other words, it simply tells one to rearrange the order of operators to a standard form, the time-ordered fonn. Thus, a product A useful symbol in mathematics is the product symbol of n functions Fl. F2, ... , F ll is often written
n.
N
(1.2)
There are many circumstances however in which 1i depends on time. A very common example arises when one is applying pulses of alternating magnetic field, HI, so that the Hamiltonian in the rotating frame makes sudden jumps from one value to another. Then during each time interval it is independent of time. During the tum-on or tum-off of the HI, which we approximate as instantaneous, the wave function does not change (a finite 1i acting for an infinitesimal time 616
-~1iltl)exp( -*1i3t3)exp( -*1i2t2)
F I F2 ··• F" '"
IT Fk
(1.9)
k"'l
Implicit in such a notation is the idea that the order of the functions, F", is not important, Le. that the Fk's commute. Thus, if we were careless we might write (J.3) as ,,(t) =
fI e.P(-ih.t.),,(O) .
(1.10)
1:=1
617
Then, if later we came to evaluate the product and continued our careless ways, we might write out the product as
,p(t) "" exp (-i1l It) exp (-i1l2t2) _.. exp (-i1l n t n)t/1(O) .
(J.1I)
If the 1lk's did not commute. we clearly would not have rewritten the correct result (l.3)! The Dyson operator allows us to deal with this problem because if we wrote (J.ll) as
,p(t) "" To exp (
-*1-h t1) ... exp (* 1l t ),p(0) n n
(J.i9)
we get the coefficient of Tn as (A2 + AI)n
(J.20)
n! Now, if At and A2 commuted, we could write (l.20) as I
we immediately get
,p(t) "" exp (
-*
n
1ln t n )ex p (
-'t (n _ '"
ltn-I) ...
-*1ln -
x exp ( -*1l2t2 )ex p ( -*1lltl ),p(0)
(J.i3)
'Therefore. we can write (l.3) compactly as
,,(t):To
IT e
•
(J.14)
I
n
_ '" n. An-k A k - n! k~ (n - k)!k! 2 I
(J.12)
I
k)!k!
A fI - k A k 2
I
which is identical to (J.18). If we expand the prodUCI of (l.20). we will get n 3 n 1 A 2n + An-IA 2 1+ An-'A 2 1 A 2+ A 2 - A I A'2'" A 2 A I A 2 - + A 1+ An-I 2 + tenns involving A 2 twice + etc (l.22) Operating on this expression by To will then give us
A n(n-t)An- 22 A n2 + n An-I 2 1+ 2 2 Al + ...
k""l
since no matter what order we use to write out the product of the exponentials. TO tells us to reorder them to get the order of (l.3). We can derive another useful theorem. Consider a product of two exponentials
(1.21)
(J.23)
which is just (J.t8). That is
TO(A, + AI)" =
(J.l 5)
nIl
L ( _ k)1 _A,-k A k
k""O n
(J.24)
. kl
We can therefore write where Al refers to some operator at an earlier time tJ, and A2 an operator at a later time t2. We can use the Dyson operator to wrile this as !(A2,AI) "" Toe Al + A\ or (J.16a) eAleA1 "" ToeAl+A\
(l.16b)
At first sight this result is surprising since (1.16) would nOl be correcl without the To present. However, the theorem is easy 10 prove by considering the related functions as power series
eA~TeA\T "" (I:I ~A~TI) (I: ~AlnTm) I. m.
(J.17)
III
To
~ eXP(*1iktk) "" To exp (* ~1iktk)
We can now deal with a general time-dependent interaction 1l(t) represented schematically by Fig. J.la. First we approximate it by a series of short steps of duration Tl,T2 • ... (Fig.J.tb) during which we approximate 1t(t) by 1t" 1t2. etc. We now use Ihis in (l.12) to wrile
,p(t) "" Toexp(* ktO 1tk Tk )J/J(O)
n
On the other hand, expanding 618
(J.i 8)
(J.26)
Nexi we let the Tk 's gel smaller, thereby increasing n, and replace the summation sign by an integral sign to get
The coefficient of Tn in the product is then
L k""O (n
(J.25)
,,(t) =
To exp
(* J
1i(T)dT
)"(0)
(J.27)
o
II is useful to keep in mind that to evaluate such an expression one can always go back to either (l.26) or indeed 10 (l.3) whenever doubts arise. 619
~ 'H Itl
10) FI,.J.\. (.) A 1,~dop""O"IIl.mi1'o,,;." 1«.) V(!rsus I (b) The Hamiltoman of (a) approxi· nlllted by a series of lIme-independenl n"millOllillns of amplitude 1ft (I: = 1 1.0 n) lind duration
-* J -* J
(a(T)"Ii. + b(T)"Ii,JdT),,(0)
"(') = Toexp(
o
= exp (
Tt
[a(T)"Ii. + b(T)"Ii,ldT) ,,(0)
o
.' =
(bl
exp{ -T.["Ii.
(J a(T)dT) + "Ii, (J, b(T)dT») ] },,(O) o
(1.34)
0
If we define the average value of a or b over the lime interval t as 'H It I
- tIf
,
- tIf
I
b(T)dT
(J.3S)
,,(t) = exp ( -*(a(')"Ii. + b(')"Ii,],) ,,(0)
(1.36)
aCt) =
'.
a(T)dT
b(t) =
,
o
Sometimes the Hamiilonian has a par1icularly simple rime dependence
(1.28)
"Ii(') = a(')"Ii.
o
then we can wrile (1.34) as
Then H(tt) and 1t(t2), the Hamiltonians al differenl times, commute
(J.29)
("Ii(',), "Ii(',)] = a(,,)a(,,)("Ii., "Ii.1 = 0 In this case
,pet) = exp ( -
*"Ii.
Ja(T)dT) ,p(0)
(1.30)
An interesting situation arises when the Hamiltonian contains several parts with different time dependences. We shall distinguish between two cases.
Case 1:
A Hamiltonian composed of several time-dependem parts which commute. We can do a similar thing if the Hamiltonian cOnlains several pans which com mUle but with differing time dependences. Lei aCt) and b(t) be two different functions of time, and let Jig and 'Hb commute: (1.31)
["Ii.,"Ii,1 = 0 Then if
(1.32)
'H(t) = a(t)H a + b(t)Hb
,
n=O,I,2,etc.
"(') = exp ( -*(a(T)"Ii. + b(T)"Ii,]nT),,(0)
(1.37)
("Ii(',), "Ii(',)] = 0(' ,)a(,,)["Ii.. "Ii. I + 0(', )b(,,)["Ii•• "Ii,] + )a(,,)("Ii,.1i.] + )b(,,)("Ii,. "Ii,) = 0
b(',
(1.33)
(J.38)
This is a fonn of stroboscopic observation. Exacdy Ihis case arises in magic angle spinning. These fonnulas are easily generalized to a Hamihonian comaining more Ihan lWO lenns a and b.
Case 11: A Hamiltonian composed of several lime-dependent parts which do 001 commute. If we have IwO functions aCt) and b(t) wilh differenl lime dependences and It Hamiltonian composed of tWO nOrlcommmirlg pans H(t) = a(t)H a
("Ii.. "Ii,]
il is slill true Ihat thai Hamiltonian at different times commules since
Therefore, we can write
t=nT
Ihen aCt) and b(t) are independent of lime and
o
b(',
In general the average values aCt) and b(t) will vary wilh t. However, somelimes Ihe lime dependence is periodic, in which case if one chooses jusl Ihose times t which are an inlegral muhiple of the basic period T,
+ b(t)'Hb
with
,,0 .
(J.39,) (1.39b)
then we have 10 use the complelely general relationship (1.27) or its equivalent fonns. If aCt) and b(t) are periodic, one can consider each period 10 define
U(T) = Toexp(
-* J
"Ii(T)dT )"(0)
(1.40)
o
620
62'
Then for stroboscopic observation
The theoretical formulas we have just given are sometimes referred to as
-k J . nT
,penT) = TDe,p (
average HamiflQnian theory. For it to hold true, typical matrix elements of 11., (al1ilo/), must satisfy
)
H(T)dT ,p(0)
o
=TDe,+~ t( k:::1
(al~la')T -<: 1
kT
J
H(T)dT)],p(O)
(k-\)T
kT
=TD
IT e,p(-~ J
H(T)dT),p(O)
(1.41)
(k-I)T
k:::l
Bol kT
J
(k-\)T
is independent of k, so
II [TD e,p(-
K. Correction Terms in Average Hamiltonian TheoryThe Magnus Expansion
*J
H( T)dT ) ] ,p(0) = U"(T),p(O)
(1.43)
o
Thus, we need only work out U(T) to find !f(nT). Suppose, then, that the time interval T is very short so that U(T) differs only slightly from the identity operator. In this case, we can keep as our lowest order approximation the first two terms in the exponential expansion of U(T), getting
.T
U(T) '" 1 - TDk
J
H(T)dT
(1.44)
o
However, thinking of the integral as a sum of infinitesimal tenos, we see that we never have a problem of commutating operators since we have no products of 11.(1') at one time with 11.(1') at another time. Consequently we can omit the time ordering operation, getting
*J . T
U(T)::: 1 -
11.(1')d1' S! exp
o
(
-* J . T
11.(1')d1'
)
(1.45)
J
11.(1')d1'
622
11. B ::: p\-l P2-111.intP2Pl
(8.152)
11.c ::: 11.int
where 11.in~ is the sum of the Zeeman and dipolar Hamiltonians in lhe rotating reference frame. Then, over one cycle, the effect of the Hamiltonian is given by the unitary operator UT of (8.153) UT ::: exp ( -*11.C1'3 )ex p ( -*11.B1'2 )ex p ( -*11.A1'l)
(8.153)
We expand the exponentials and keep just the leading terms, to write
::' ex p ( -*(11.C1'3+11.B1'2+11.A1'd) (1.46)
The approximation is equivalent to writing eCebea = ec+b+ a
we can write
e,p( -*HT)
::: p\-I1iintPt
r.
o
U(T) =
1i A
UT :::1--(11.CT3+11.BT2'HA1'l)
1 T
T
In the theory of spin-flip line narrowing as well as in multiquantum excitation the effect of applications of sequences of pulses is analyzed using a time-averaged Hamiltonian. Specifically, in Sect. 8.11 we analyze a three-pulse sequence which causes the Hamiltonian to be effectively transformed as described in (8.152):
;
•
Defining the time average of the Hamiltonian 11. as 11.:::
One useful way to think of this approximation is in terms of rotations. Rotations in three dimension do not commute, but infinitesimal ones do. Thus, if we think of 'H a and 11.b as two rotations, their total effect is changed if we reverse the order in which we perform the rotations. If, however, the rotations are infinitesimal, corresponding to satisfying (J,48), the order of performing the rotations is irrelevant. In Appendix K we take up the corrections to the formula of (J,47).
(1.42)
11.(1')dT
,p(nT) =
(1.48)
(1.47)
(KI)
for three nOllcommuting operators a, b, c. Since clearly the expression (KI) is only approximate, one would like to know how big the errors are. This problem 623
has been treated by many authors. A pioneering paper in this regard is the treatment of Haeberlen and Waugh [K.IJ. A useful mathematical treatment is that of Wi/cox [K.2]. The basic paper is that of Magnus [K3], who derived the first terms of a series expansion for a time-dependent Hamiltonian 1{(t) as being equivalent to a unitary operator Vet) given by
U(t) = exp ( -*('Ii + ?i(l) + 'Ii(2»t) where
(K.2)
,
- tIJ 1{ =
(K.3a)
1{(tddt l
o .
t
t1
2:. Jo J tit,
'Ii(l) = -
tit, ['Ii(t,),'Ii(t,)]
(K.3b)
0
It t3/1 'Ii(') = - , dt, tit, dtt!['Ii(t,), ['Ii(t,),?i(tl)J) 6th 0 0 0 + ['Ii(t,), ['Ii(t,),'Ii(t,»)J)
J J J
(K.4) =
(1 +e+ ;~ + .. .) (I +b+ ~~ + .. .) (I +a+ ~~ + .. .) !
= 1 + (c + b + a) + 2! (c 2 + b2 + a2 + 2eb + 2ea + 2ba) + ...
(K,5)
where in (K,4) and (K,5) we have collected all the terms involving products of 0, I, or 2 of the operators but have not listed those involving 3 or more. Comparing, we see that (K.4) contains terms such as bc+eb, whereas (K,5) has only 2cb, i.e. e only to the left of b. We can therefore write eCebe tl
';;!
'"
624
dtl1i{tl) =
0
f-c [i
dtl1i(t.) +
0
J
dtl1i(td+
tA
J
to
!
= -[1i A TA + 1iBTB + 1icTcl
te
dt l1i(tl)] (K.7)
This is simply the average Hamiltonian we have used previously. We turn now to the next term, 1i(l), given by (K.3b):
J, dt, J" dt!l'li(t,),?i(t,)] o
I
a
f-c J
'liP! = -2:1,
= I + (c + b + a) + 2! (c 2 + b2 + a2 + cb + be + ca + ac + ba + ab) + ...
(K.6)
where in these expressions there is implicit the idea that the expression is to be thought of as an expansion, and is accurate only up to terms involving products of pairs of the operators. As we show below, the expressions (K.6) is an example of including just the first two terms, 1{ and ?t(1), of the Magnus expansion. Deriving the result (K,6) was easy. It is obvious now how to get the next term for our simple example of (K.I): merely by continuing the expansion of the exponentials. We leave that as an exercise for the reader. We now want to show that if we evaluate (K.3), we get the same result as (K.6). First let us define times tA, tB, and tc as follows. The time-dependent Hamiltonian exists as 1i A from t = 0 to t = tA, as 1iB from tA to tB, and as 1ic from tB to tc. We further define TA' TB' and TC as the durations of the three.intervals. Then we have, for one cycle, from (K.3)
(K.3c)
I ec+b+a = I +(c+b+a)+ 2t(c+b+a)2+ ...
eCebe
exp (e + b+ a)exp [!([c, bJ + [e, a] + [b, a])J
'" exp [(c+ b+ a) + ~(Ie, h] + Ie, aJ + [b, aD]
1i =
Our goal is to give the reader a simple idea of where these expressions come from, and what they say in concrete terms. To do so, let us consider the three pulse example. Let us therefore examine the two sides of (KI), utilizing power series expansion of the exponentials. Thus
whereas
';;t
(K.3b)
0
Now t2 is always later than or equal to t 1. Thus, they could both fall into the same time interval, such as between 0 and t A , or t2 could be in one interval (e.g. between tA and tB) and t1 in another (in this case necessarily 0 to tA)' Since 1i(t) is a constant during anyone interval, it is useful to break the integrals into these time domains. We therefore adopt a noration for specifying the limits of integration: tc
t1
o
0
Jdt2 Jdtl F(t2,
tl)
{exp(e+b+a)} +!(eb-be+ea-ae+ba-ab)+ ... exp (e + b- a){1 + ~([e, b] + [e, aJ + [b, aD)
(K.8)
625
Consider the first integral. Since tl and t2 are both between 0 and tA, L....
,2:SL....
o
0
I....
An imponam application of these equations is to find pulse sequences which cause 11.(1) to vanish, Here is a simple example. Suppose that one considers a pulse sequence of equal T'S. Then
h :S t ....
Jd', J d', [H(.,). H(.,)] : Jo d', J d.,[HA.HA] (K.9)
:0
Consider then the second tenn on the right side of (K.8) to
t....
In
0
L....
I....
(K.IO)
te
t2
0
0
2'C"
(K.ll)
so that using (K.ll) the term of (K.2) involving 1-£(1). -i1f!..l)tc/h, is
2"
+ [Hc.HAITcTA + [Hc.HBITcTB) i
b= --",'HBT8
(K.12)
i c = --'HcTc
"
(K.13)
1
1
2" which agrees with (K.12).
+ ['HC.11. A]TCTA + ['HB, 11.",]TSTA),
(K. 14)
It is clear that if one has a succession of intervals j = I to N. of duration and Hamiltonians 11.j, the general form of 'H(I) is
-2~h ,2:
('Hj.'Hk]TjTk
where
(K,ll.)
T:
I>; j=1
625
true
for the Hamiltonian of (8.158). Then, since
[H"HtI:O
.
Alas. this does nOI vanish. How can we construct a pulse sequence which vanishes? Haeberlen and Waugh point out that one can readily find a six-pulse sequence which vanishes. Introducing the notation
[H;.H,I " i. k
(K,21)
S = 6,5 + 6,4 + 6,3 + 6,2 +6, 1 +5,4+5,3+5,2+5,1 +4,3+4,2+4,1 +3,2+3,1 + 2,1 Now, previously we had 11. 1 + 11.2 + 'H 3 = 0
(K,22) (K.23)
Suppose we keep this condition, and add to it the condition
J>k
N
as was
(K.19)
we have
2([e. b] + [c. al + [b. al) : --.2 ([1t'C.'HB]rC TB
(K.18)
(K.20)
we get that the correction tenn of (K.6)
'H(I) =
For a three-pulse sequence this can be written
1 = - - , (['Hs, 1{",)TATS
'
(K.17)
we can write
Now. identifying a. b. and c as
Tj
[H;.H,I .
1-£1+1-£2+1-£3=0
: _ _i_({'H8, 'H",lT8TA + ['Hc, 1t'",]TCTA + (1{c, 1(8)TC1'"S)
i a= --",11. AT",
(K.16)
Now suppose
Jd', Jd'J!H(.,).H(tI)] 2tch
Jill) : - - ' -
"
L
S: [H,.H,I + [H,.HtI + [H,.HtI .
i'H(1lt c
[H;.H,)
j>k
We therefore get
-
L
2Th j >k
S:
0
= [1i8' 'HA)T8 TA
,
H(I): -~
so that the vanishing of 11.(1) arises from the vanishing of the sum S defined as
Jd', Jd', [H(.,).H('I)] : Jd', Jd'I[HB.HA]
I....
.
.
0
(K.Ilb)
'H4+'HS+11.6=O
.
(K.24)
Then, adding the columns of (K.22) venically, and insening items such as 4,4 in the second column, etc. we get that the second, third, and fifth columns add to zero, leaving solely contributions from the first and founh columns: 627
s =6, 5 + 3,2 =['Ii"
Selected Bibliography
'Ii,] + ['Ii" 'li2]
It is now obvious that if we pick
1i6 '" 11.2
1£5 '" 113
the commutators will be each other's negative. with the result that S will vanish. Thus. a cycle in which (K.23) is [flIe and in addition ~=1iA=~
~=1iB=~
~=~=1il
will have a vanishing 1-£(1).
There is an ex.tensive literature on methods of finding such cycles. See for example [K.4,5J.
The problem of preparin& .. complete bibliO&~.lIphy of magnetic re!lOlilInce is hopeless, there have been so ffiilny papers. Such .. bibliography would not even be usdul for 11 student, since he would not know whcre to begin. Therdore, a short list of articles has b~n selected which touches on II number of the most important ideas in resonl'lncc. In some instances, papers were chosen because they are basic references; others because they were representative of a dl\SS of papers. In some CMCll 1\11 aLtempt hIlS been made Lo augment the treatment of the text. An important technique for sear<:hing the literatu~ ill use of the Citation Index, which lists all the papers in any given year which refer to a particular article or book (r-pers in the Citation Index are listed by the first name among the authors, e.g. by B~mbergen for the paper by Bioembergen, Purcell, and Pound). ThUll to search for current work in • given area, one can look up • basic earlier paper to which later authors are likely to refer. The use of the Citation Index in connection with the following bibliography is the best method of obu.ining an up-to-date picture of ....-ork in the areas listed.
Basic Papers E.M. Purcell, II.C. Torrey, R.V. Pound: Re!IOnaoce absorplion by nuc:lear magnetic mo-ment.8 in atolid. Phys. Rev. 69, 37 (1946) F. Bloch, W.W. Hansen, M. Paek&rd: Nuc:lear inductH>n. Phys. Rev. 69,127 (1946) F. Bloch, W.W. Hansen, M. Paebro: The nuc:lu.r inductH>n experiment. PhYll. Rev. 70, 474-485 (1946) F. Bloch: Nuc:lear induction. Phys. Rev. 70, 460-474 (1946) N. Bloembergen, E.M. Purcell, R.V. Pound: Relaxation effects in Iluc:lear magnetic res0nance absorption. Phys. Rev. 73, 679-712 (1948) Boo~s,
Monographs, and Review Articles
G.E. Pake: -Nuelear Magnetic Resonance·, in Solid Slale PlIpics, Vol. 2, ed. by F. Seitz, D. Thmbull (Academic, New York 1956) pp. 1-91 Collected artic:les. Nuovo cimento Suppl., Vol. VI, Ser X, p 808 fr. (1957) E.R. Andrew: Nuclear MagMI~ ReSOMfll:e (Cambridge University Press, Cambridge 1955) J.A. Pople, W.G. Schneider, II.J. Bernstein: lligh·Resolulitm Nudear Magll£lic ResoNlfll:e (McGraw-llill, New York 1959) C.J. Gorter raratllagnel~RtI='ion(Elsevier, New York 1947) O.J .E. Ingram: Spectroscopy al Radio and Microwave Frequencies (Butterworth, London 1955) A.K. Saha, T.P. Oas: Theory and A-pplit:alions of Nudear Induclion (Saha Inslilule of Nuclear Phyllics, Calcutta 1957) M.II. Cohen. F. Reif: ~uadrtJpole ER'ccl.s in Nuclear Magnd.ic Resonance Sludies of Solids", in Solid Stole Physics, Vol. 5, ed. by F. Seitz, D. Turnbull (Academic, New York 1957) pp. 321-438 T .1'. Oas, E.L. lIahn: "Nuclear Quadrupole Resonance Spedroscopy· , in SoIidSkilt Physics. Supplement 1, ed. by F. Seitz, D. Thrnbull (Academic, New York 1958) T.J. Rowland: Nuclear magnetic resonance in metals. Prog. Maltr. Sci. 9,1-91 (1961)
628
828
II .12 Coupling
R.V. Pound: Prog.Nucl. Phys. 2, 21-50 (1952) . William Low: ~Paramagnclic Resonance in Solids", in Solid Start PhysICS, Supplement 2, ed. by F. !kit~, D. Turnbull (Academic, New York .1960). . . J.S. Griffith: TM Thea? ofTr(lllSilUm·Mefalloru (Cambridge Umvennly Press, Cambndge 1961) N.F. Ramsey: NucltoT MonunlS (Wiley, New York 1953) N.F. Ramsey: Moiteular Beams (Clarendon, Odord 1956) A. Abragam: The Principles ofNudear Magnefism (Clarendon, Ox£O«I 1961) N. Bloembergen: NuclearMag~f~Re/arJJlUm(W.A. ~njamin, New York 1961) J.D. RoberLs: Nuclear Magntl~ ReSDfl(JllCe (McGraw-lilli, ~ew .York 195~) R.T. Schumacher: IntroducliOfl 10 Magntlic RtSOfI(J)lU (Bel\lamm-Cummmgs, Menlo Park, CA 1970) I{ .A. McLauchll\ll: Magnetic Resonance (Clllrcndon, Oxford 1972) . . G.E. Pake, T.L. Estle: TM Physical PrillCiples of Electron Paramllgnttic ResontJflCe (BcnJammCummings, Menlo Park, CA 1973) .. A. Abragam, B. Sleane)': Elecfron Paramagntlk Resonat1Ce o/TrarulflQn Ions (Clarendon, Odord 1970) . Maurice Goldman: Spill Tentplral/Ut tlfId Nuelear Magntlk RtuJNUlCe in Solids (Clarendon, Oxford 1970) . ' C. Kittel: lnlroductiOfiIO Solid SIo4 PhyslU, 5th ed. (WIley, New Y~rk 1976) Chap. 16 C.P. Poole, Jr., II.A. Faraeh: TM TMory c{Magntlic RUOIVUtCe (WIley, New York 1972)
11.5. Gutowsl:y, D.W. McCall, C.P. Slichler: Nuclear magnelic fCIIOnance mulliplels in liquids. J. Chem. Phys. 21, 279-292 (1953) E.L. Hahn, D.E. Maxwell: Spin echo measUT(!menls of nuclear spin coupling in molecules. Phys. Rev. 88, 107G-108-I (1952) N.F. Ramsey, E.M. Purttll: InteractiOn5 beh.:ccn nuclear spins in molecules. Phys. Rev. 85,1>\3-1>\4 (1952) (Idter) N.F. Ramsey: Electron coupled illleracl;ons between nuclear spins in molecules. I'hy•. Rev. 91, 303-307 (1953) N. Bloembergell, T.J. Rowland: Nudearspin exchange in solids: 1'1103 and 1'1 105 lllll.gnetic resonance in thalliu.,n and ll~allic oxide. Phys. Rev. 97,1679-1698 (1955) M.A. Ruderman, C. I{,ttel: Indirect exchange coupling of nuclear magnetic momenLS by conduclion eleclrons. Ph)·s. Rev. 96, 99-102 (1954) K. Yosida: ~lagnetic properties of CIl-Mn alloys. Phys. Rev. 106, 893-898 (1957) H.M. McC~lllnell, A.D. McLean, C.A. Reill)': t\nalysis of spin-spin multiplet.s in nude"r magnetic reI:IOnance speclra. J. Chem. Phys. 23, 1152-1159 (1955) II.M. McConnell: Molecular orbilal approximll.lion lo e1eclron coupled inlerAction between nuclear spins. J. Chem. Phy•. 14, 46G-467 (1956) W.A. Anderson: Nuclear magnelic resonance speclra of some hydrocarbons. I'hy•. !lev. 102,151-167 (1956)
General Theory or Resonance
Pulse Methods
D. Pines, C.P. Slichler: RelaxaLion limtlJ in magnetic resonance. Phys. Rev. 100, 10141020 (1955) f1.C. Torrey: Bloch equations wilh diffusion terms. Ph)'s. Rev. 104, 563-565 (1956) R. Kubo, K. Tomila: A general theory of magnetic resonance absorption. J. Phy!. Soc. Jpn. 9,888-919 (1954) P.\V. Anderson. P.R. Weiss: Exchange narrowing in paramagnetic resonance. Rev. Mod. Phys. 2S, 209·276 (1953) . ' P. W. Anderson: A mathematical model for lhe narrowmg of spectral hnes by exchange or motion. J. Phys. Soc. Jpn. 9, 316-339 (195)\) .. n.K. Wangness, F. Bloch: The dynamical theory of nuclear mductlon. Phy•. Rev. 89, 72S-739 (1953) F. Bloch: Dynamical theory of nuclear induction II. Phy•. Rev. 102, 104-135 (1956) A.G. Redfield: On the theory of relaxation processes. IBM J. I, 19-31 (1957) II.C. Torrey: Nuclear spin relaxat;on by translational diffusion. Phys. Rev. 91, 962-969 (1953)
E.L. f1ahn: Spin echoCll. Phys. Rev. 80, 58G-59>\ (1950) II.Y. Carr, E.M. Purcell: Effects of diffusion on free precession in nude..r magnelic resonance experiments. Phys. Rev. 94, 63G-638 (1954)
Nuclear Magnetic Resonance in Metals C.II. Townes, C. Herring, W.O. Knight: The effect of electronic paramagnetism on nuclear magnetic resonance frequencies in metals. Phy•. Rev. 77, ~52-853 (1950.) (ldter). . W.O. Knighl: ~Eleclron Paramagnelism and Nude...r Magnetic Rt$Qnance ill Mdall ,In SolidS/alt Physics, Vol. 2, ed. by F. Seilz, D. Thrnbull (Academic, New York 1956) pp. 93-136 J. Korringa: Nudear magnelic relaxation and resonance line shifl in metals. Physical6, 601-610 (1950) D.F. Holcomb, R.E. Norberg: Nudear spin relaxation in alkali metals. Phys. Rev. 98, 1074-1091 (1955) . G. Benedek, T. Kushida: The pressure dependence of the Knighl shirt in the alkah metal. and copper. J. Phys. Chern. Solids S, 241 (1958)
630
Second Moment L.J.F. Broer: On the tl~oory of param~gnelic relaxation. Physiea 10, 801-816 (1943) J.II. Van Vleck: The dipolar broadenlllg of magnetic resonance lines in crystals. Phys. Rev. 74, ]]68-1183 (1948) G.E. Pake: Nuclear resonance absorption in hydrated cry.Lals: Fine struclure of lhe proton line. J. Chem. Ph)'s. J6, 327_336 (1948) 11.5. GUlowsky, G.E. Pake: Nudear magnetism in studielll of molecular struc:tUrt and rotatx)fl in solids: ammonium sal~. J. Chem. Ph)'s. J6, 1l64-1165 (l!H8) (letter) E.R. Andrew, R.G. Eades: A nuclear magnelic resonance investigation of lhree IIOlid benzel1t$. Proc. R. Soc. London A118, 537-552 (1953) H.S. G~towsky, G.E. Pake: Structural investigat;ons by means of nudear magnetism II - IImdered rotalion in solids. J. Chem. Phys. 18, 162-170 (1950)
Nuclear Polarization A.W. Overhauser: Polari;/;alion of nudei in metals. Phys. Rcv. 91, 411-415 (1953) T.R. Carver, C.P. Slichlcr: Experimental verification of lhe O"erhauser nudear po[ari;/;alion elfOX:l. Phys. Rev. 101,975-980 (1956) A. Abraga.m: Overh.au8l?r effecl in n.onmetals. Ph)'s. Rev. 98, 1729·1735 (1955) C.D. Jeffnes: PolarlzallOn of nuclei by resonance saturation in l:>aramagnetic crystals. Phys. Rev. 106, 164-165 (1957) (Ieller) J. U.e~feld, J.L. MOl.chane, E.Erb: Augmentation de II. polarisation nucleaire da", les hqmdes el gu adsorbes sur un charbon. Extension aux solidc8 conlenanl des impuretes paramagnetiques. J. Ph)'ll. Radium 19, 843-84'4 (1958) A. Abr~am, W.G. Proct.or: Une nouvelle melhode de polarisation dynamiquedes noyaux atomique dan. Its 8Olides. C.R. Acad. Sci 146, 2253-2256 (1958)
631
C.D. Jeffries: "Dynamic Nuelear Polarization- in Progrusin Cryo~nia (Heywood, London 1961) G.R. Khulsishvili: The Overhauser effed and related phenomena. Soviet Phys. - Usp. 3, 285-319 (1960) R.H. \Vebb: Steady-state nuclear polarizations via electronic transitions. Am. J. Phys. 19,428-444 (1961)
Quadrupole Effects R. V. Pound: Nuclear electric quadrupole inleractions in crystals. Phys. Rev. 79, 68:>-702 ( 1950) N. B1ocmber8en: "Nuclear Magnetic Resonance in Imperfect Crystals", in Report of the Bristol Conference on Defects in Crystalline Solids (Physical Society, London 1955) l>p. 1-32 T.J. Rowland: Nuclear magnetic resonance in copper alloys. Electron distribution around mlute atoms. Phys. Rev. 119, 900-912 (1960) W. Kohn, 5.11. Voslco: Theory or nuclear rCllOnance intensity in dilute alloys. Phy!l. Rev. 119,912-918 (1960) T.P. Das. M. Pomeralllz: Nuclear quadrupole interaction in pure metals. Phys. Rev. 113, 2070 (1961) T. Kushida, G. Benedek, N. Bloernbergen: Dependence of pure quadrupole resonance frequency on pressure and temperalure. Phys. Rev. 1G4, 1364 (1956)
Chemical Shifts W.G. Proctor, f.C. Yu: The dependence of a nuclear magnetic resonance frequency upon chemical compound. Phys. Rev. 77 , 717 (1950) W.C. Dickinson: Dependence of the piS nuclear resonance position on chemical compound. Rev. 77, 736 (1950) 11.5. Gutowsky, C.J. 1I0ffmann: Chemical shins in the mllgnetic I'ClIOnance of F1S. Phys. Rev. 80, 1I0-lll (1950) (letter) N.F. Ramsey: Magnetic: shielding of nuclei in molecules. Phys. Rev. 78, 699-703 (1950) N.F. Ramsey: Chernic:a1 effect.s in nuclear rnasnetic resonance and in diamasnetic susceptibility. Phy!l. Rev. 86, 243-246 (1952) A. Saika, C.P. Slichter: A note on the nuorine I'ClIOnance shifts. J. Chern. Phya. 22, 26-28 (1954) J.A. Pople: The theory of chernielll llhms in nuclear magnetic resonance I - Induced current denllities. Proc. R. Soc. London A139, 541-549 (1957) J .A. Pople: The thoory of chemicalshifb in nuclear magnetic resonance II -Interpretation of proton shiftll. Prot. R. Soc. London A1J9, 550-556 (1957) II.M. McConnell: Theory of nuclear magnetic shielding in molecules 1- Long range dipolar shielding or protons. J. Chern. Phys. 17, 226-229 (1957) R. ~man, G. MUrTay, R. Richards; Cobah nuclear resonance llpect.... Proe. R. Soc. London 141A, 455 (1957)
Spin Temperature N. Bloembergen: On the interaction of nuclear apins in a crystalline lattice. Phyaica 15, 386-426 (19-49) E.M. Purcell, R. V. Pound: A nuclear spin llyalem at nqative lemperatu~. Phy•. Rev. 81,279-280 (1951) (Iella-) A. Abragarn, W.G. Proctor: Experiment!l on spin temperature. Phys. Rev. 106, 160-161 (1957) (letter) A. Abragam, W.G. Proctor: Spin temperatUl"e. PhYll. Rev. 109, 1441-1458 (1958)
A.G. Redfield: Nuclear magnetic resonance saturation and rotary saturation in solidll. Phys. Rev. 98, 1787-1809 (1955) A.G. Redfield: Nuclear spin·lallice relaxa~ion time in copper and aluminium. Phys. Rev 101,67·68 (1956) . J.II. Van Vlcek: The physical meaning of adiabatic magnetic llU!ICep~ibilities. Z. PhYll. Chern. Neue Folge 16, 358 (1958) J.II. Van Vleck: The coneept oftemperature in magnetillm. II Nouvo Cimento Suppl 6 Serie X, 1081 (1957) , ., C.P. Stichter, W.C. 1I01ton: Adiabatic demagnetization in a rotating reference llystem. PhYll. Rev. 121, 1701-1708 (1961) A.G. Anderson, A.G. Redfield: Nuclear spin-laUice relaxation in metals PhYll. Rev. 116, 583-591 (1959) . L.C..H~bel, C.P. Slichter: Nuclear llpin relaxatiOll ill Normal and SIII)erconduc~il\g alumllllUm. PhYll. Rev. 113, 1504-1519 (1959) A. Andcrson: Nonresonant nuclear spin absorption in Li Na and AI. PI,Yll Rev liS 863 (1959) , , ", L.C. liebel, Jr.: Spin Temperature and Nuc:lcar Relaxation in Solidll Solid SWlt Physics Vol. 15 (Academic, New York 1963) "
Rate Effects H.S. GutO\':'llky, A. Saika: Dissocia~ion, chemical exchange, and the proton magnetic resonance In lOme '"'!uoous electrolytes. J. Chelll. PhYll.l1, 1688-1694 (1953) J.T. Arnold: MagnetiC re!IOnancea of protollll in ethyl alcohol. PhYll Rev 102 136-150 (1956) . . , 9 R. Kuoo: Note on the lltochastic thoory or resonance abl'lorption. J. Phys. Soc J . pn. , 935-944 (195-4) U.M. McConnell: Reaction ra~es by nudear magnetic I'ClIOnanee. J. Chem Phys 18 430431 (1958) . " S. Meiboom. Z. I,uz, D. Gill: Proton relaxation in water. J. Chelll. Phys.17 1411-1412 , (1957) (letter)
Cross·Relaxation N. Dloembergen, S. Shapiro, P.S. Pershan, J.O. Artman: Cr06S-reiaxation in llpin SYlltems Phys. Rev. 114, 445-459 (1959) . P.S. Pershan: Cross relaxation ill tiF. Phys. Rev. 117, 109·116 (1960)
Electron Spin Resonance in Paramagnetic Systems D. Dleaney, K.W.f1. Stevens: Paramagnetic resonance. Rep. Prog. Phys. XVI, 108-159 (1953) T.G. CBlItner, W. Kanzig: The electronic structure of V-centera. J. Phys Chem Solids 3,178-195 (1957) . . . G.D. Wat.kinll: EI.eclro~ ~pill resonallee of Mn++ in alkali chloridell: association with vacancl,:" and IIl~PUfltll:S. PhYll. Rev. 113, 79-90 (1959) G.D. Watkms: Mo t.IOn of. Mn++-cation vacancy pairs in NaC1: !ltudy by eloctron spin resonance an d d ~eIceltiC loss. Phya. Rev. 113,91-97 (1959) G. ~eher: Ob6ervahon of nuclear masnetic resonances via the electron spin re!IOnance Ime. PhYll. Rev. 103, 83-4-835 (1956) G. Feher: Electronic structure or F centera in KCI by the electron llpln double resonance technique. PhYll. Rev. 105, 1122-1123 (1957) G. Feher: Electron llpin I'ClIOnance experiments on donofll in llilicon _ Elec~ronic struelure of donors by the electron nuclear double reSOllallce technique. Phys Rev. 114 1219. 1244 (1959) . ,
632 633
11.11. Woodbur)', C.W. Ludwig: Spin resonance of transition metals in silicon. Phys. Rev. 117,102-108 (1960) A.F. Kip, C. Kittel, R.A. Levy, A.M. Portis: Electronic structure of F centers.: hyperline interactions in eledron spin resonance. Phya. Rev. 91,1066-1011 (1953) C.J. Delbecq, B. Smaller, P.II. Yuster: Optkal absorptioll ofCh molecule-ions in irradi_ ated potassium chloride. Phys. Rev. III, 1235-1240 (1958) M. Weger: Passage dfects in paramagnetic resonance experiments. Bell 5yst. Tech. J. 39, 1013-1112 (1960) (Monograph 366J) George Feher, A.F. Kip: Electron spin resonance absorption in metals I - Experimental. Phys. Rev. 98, 337-348 (1955)
Nuclear Resonance in Ferromagnels A.M. Portis, A.C. Gossard: Nudear resonance in ferromagnetic cobalt. J. Appl. Phys. 31, 2055-2135 (1960) W. Marshall: Orientation of nuclei in ferromagnet!>. Phys. Rev. 110, 1280-1285 (1958) R.E. Walson, A.J. Freeman: Origin or effec~ive fields in magnetic materials. Phys. Rev. 123, 2027-20H (1961) G. Benedek, J. Arms~rong: The pressure and ~emperature dependence of the FeH nuclear magnetic resonance frequency in ferromagnetic iron. J. Appl. Phys. 32, IOG5 (1961) P. Ileller, G. Benedek: Nudear magnetic resonance in MnFt neu theeritic&l point. Phys. Rev. Leu. I, 428 (1962)
Nuclear Resonance in Paramagnetic and Anliferromagnelic Substances
Double Resonance in Liquids F. BIocb: Recent developments in nuclear induction. Phys. Rev. 93, 944 (1954) V. Royden: Measurement of the spin snd gyromagnetic ratio o( Cl3 by the collapse of spin-spin splitting. Phys. Rev. 96, 543 (1954) A.L. Bloom. J.N. Schoolery: Effects of perturbing radio(requency fields on nuclear spin coupling. Phys. Rev. 97,1261 (1955) W.P. Aue, E. Bartholdi, R.1l. Ernst: Two-dimensional spectro!:lCOPY: application to nuclear magnetic resonance. J. Chern. Phys. 64, 2299 (1976)
Masers and Lasers J.P. Gordon. II.J. Zeiger, C.II. Townes: Molecular microwave 05Citlator and new hyperfine structure in the microwave spectrum of NH 3 . Phys. Rev. 95, 282 (1954) N. Bloembergen; Proposal (or a new type solid state mll8
Slow Motion C.P. Slichter, D.C. Ailion: Low-field relaxation and the study o( ultrll8low atomic motions by msgnetic resonsnce. Phy•. Rev. 135, AI099 (1964) D.C. Ailion, C.P. Slichter: Ob6ervation of ultra-slow translational diffusion in metallic lithium by magnetic resonance. Phys. Rev. 1)7, A235 (1965) D.C. Look, I.J. Lowe: Nuclear magne~ic dipole-dipole relaxation along the static and rotating magnetic fields: application to gypsum. J. Chern. Phys44, 2995 (1966)
N. Bloembcrgen: Fine structure or the magnetic resonance line or protons in CUS04 • 511 20. Physica 16, 95 (1950) N.J. Poulis, G.E.G. lIandeman: The temperature dependence of the spontaneous magnetization in an sntiferromagnetic single crysLaI. Physica 19, 391 (1953) R.G. Shulman, V. Jaccarino: Nuclear magnetic resonance in paramagnetic MnF. Phys. Rev. 108, 1219 (1957) N. Jaccarino, R.C. Shulman: Observation of nuclear magnetic resonance in anti(erromagnetic MnF, Phys. Rev. 107, 1196 (1957) G. Benedek, T. Kushida: Nuclear magnetic resonance ill antiferromagnetic MnF2 under hydrostatic p~ure, Phys. Rev. 113, 46 (1960) \V. Marshall, R.N. S~uart: Theory of trallsition ion complexes. Phys. Rev. 123, 2048 (1961)
G. Feher: Electronic structure o( F cente... in KCI by the electron spin double resonance technique. Phys. Rev. 105, 1122 (1957) II. Seidel, H.C. Wolf: In Physic.sc{ColorCefllers, 00. by W. Beall Fowler (Academic, New York 1968)
Ferromagnetic Resonance
Spin.F1ip Narrowing
B. Lax, K. Button: MicrrJWQve Furilts fJIVI FurilMlnels (McGraw-lIi11, New York 1962)
Double Resonance in Solids S.lt. Ibrtmann, E.L. Hahn: Nuclear double resonance in the rota~ing frame. Phys. Rev. 128,2042 (1962) F.M. Lurie, C.P. Slichter: Spin ~mlX'rature in nuclear double resonance. Phys. Rev. 1)3, Alt08 (1964) R.E. Slusher, E.L. lIahn: Sensitive detection o( nuclear quadrupole interactions in solKb. Phy•. Rev. 166, 332 (1968) A.G. Redfield: Pure nuclear electric quadrupole resonance in impure copper. Phys. Rev. 130,589 (t963) D.E. Kaplan, E.L. lIahn: Experience de double irradiation en resonance magnetique par la methode d'impulsions. J. Phy•. Radium 19, 821 (1958)
634
ENDOR
J.S. Waugh, L.M. Huber, U. Ilaeberlen: Approach to high-resolution NMR in solids. Phy•. Rev. 20, 180 (1968) U. llaeberlen, J.5. Waugh: Coherent averaging effect. in magnetic resonance. Phys. Rev. 175,453 (1968) P. Mansfield: Symmetrized pulse sequences in high resolu~ion NMR in solids. J. Phys. C4, 1444 (1971) P. Mansfield, M.J. Orchard, D.C. Stalker, K.H.B. Richards: Symrne~rized mul~ipulse nuclear-magnetic-resonance experiments in solids: mell.llurernent o( the chemical shirt ahield~ng tensor in some compounds. Phys. Rev. 87, 90 (1973) W.K. Rhlm, D.O. EUeman, R.W. Vaughan: Analysis o( multiple pulse: NMR in 8Olids. J. Chern. Phy•. 59, 3740 (1973) W.K. Rhim, D.O. Dleman, L.B. Schreiber, R.W. Vaushan: Analysis o( multiple pulse NMR in solids. II. J. Chern. Phys. 60, 4595 (1974) M. Mehring: 1/igh ResdUfUM NMR SpeetTOSCOf11 j"SoIids, NM R: Basic Principles and Progress, Vol. II, ed. by P. Diehl, E. Fluck, P. KOlIfeld (Springer, Berlin, lIeidelberg 1976)
635
U. Ilaeberlen: lIi,h Ra
E.O. Stejskal, J. Schaefer, R.A. McKsy: Iligh-resolution, slow-spinning magic-angle carbon_ 13 NMR. J. ~I&&n. 11.eson. 25, 569 (1917) W.T. Dixon: Spinnin,-.ideband-free and spinning.sideband-only NMR !lpec:tra in spillni~g samples. J. Chern. Phys. n, 1800 (1982)
Chemically Induced Nuclear Polarization (CIDNP) G.L. Closs; A mechanism explaining nuclear spin polaritations in radical combinalion reactions. J. Am. Chern. Soc. 91, ..552 (1969) . . G.L. Ooss, A.D. Trifunac: Chemically illd~ed n",?lear sp," pohlrlZllllon Illl a tool for determination of spin multiplicities of radical-pair precursors. J. Am. Chern. Soc. 91, . . I ' l" 4554 (1969) J.II. Freed, J.B. Pederson: The lheory of chemically induced dynamIc Spill po arlza Ion. Adv. Magn. 11.C!lOn. 8, 2 (1976) .. C.L. Closs: Chemically induced nuclear polarlzallon. Adv. Magn. Reson. 7,157 (HI74)
Composite Pulses M.II. Levitt, R. F'r-eeman: NMR population inversion using a composite pulse. J. Magn. Reson. 33, 473 (1979) M.II. Levitt: Composiu pulses. Pros. NMR S~t~: 18, 61. (1~86,> A.J. Shaka, J. Keeler: Broadband spin decouphng III IlIOtroPIC liqUids. Prog. NMR Spectros. 19, 47 (1987) . . h' h i ' I M.II. Levitt, 11.. Freeman, T. Frankie!: Broad band decouplml III '15 reso utlon nuc ('ar msgnetic reso;lsllCe !lpedroscopy. Adv. Magn. Reson. 11, 48 (1983)
Electron Spin Echo Modulation L.G. Rowan, t~.L. JI"hn, W.B. Mims: Electron !lpin-echo envelope modulation. Phys. Rev. 137 A61 (1965), errata Phys. Rev. 138, AB4 (196~) W.B. Mims: Envelope modulation in spin-echo experIments. Phys. Rev. 85, 2409 (1972)
Fourier Transrorm NMR I.J. Lowe, R.E. Norbe,"&: Free-induction de<.ays.in 9Olids.. P~ys. Rev. 107,"6 (1957) C.R. Bruce: Fit nuclear magnetie resonance hne shapcs III CaF,. Phys. Rev. 107, 43 (1957) . . 11..11.. Ernst, \Y .A. Anderson: Applications of Founer transform spectroscopy to malnehc resonanc('. ReY. Sci. Instrum. 37, 93 (1966) . ) ED Ikcker: puJu. and F~r TrlUlSform NMR (AcademIC, New York 1~87 T .C . F&rrsr, . . . ' __ .J N ,- M lie RtsOMIICe III OM R.R. Ernsl, G. Bodenhausen, A. Woksun: PrillClP~", ~Kor agM aM Tl\AO DimellSWns (Clarendon, Oxford 1987)
Magic Angle Spinning
Magnetic Resonance Imaging P.C. wuterbur: Image formation by induced local interactions: Exsmple!l employing nu· clear magnetic resonance. Nature 141,190 (1973) P. Mansfield, P.K. Granell: NMR "dilTractiOIl~ in $()lids? J. Phys. C6, L422 (1973) A. I{umar, D. Welli, Il.R. Ernst: NMIl Fourier zeugmalography. J. Magn. Reson. 18, 69 (1975) K.K. King. P.R. Moran: A unified description of NMR iml\ging data-colleclion strategies, snd reconstruction. Med. Phys. II, 1 (1984) P. Mansfield, P.G. Morris: NMR IrMgiflll ill DiomediciM, ed. by J.S. Waugh (Acsdemic, New York 1982)
Superoperators J. Jeenes-: Supero~"'tors in magnetie resonance. Adv. Ma~n. Reson. 10,2 (1982) R. Zwanzig: Ensemble method in theory of irreversibility. J. Chern. Phys.33, 1338 (1960)
Some Recent Books or Special Interest F.11.. Bovey: Nw;fear MagMlic RuorJal1Ce Spectroscopy (Academic, New York 1988) \V.S. Brey: Pulse Methods in lD aNl2D Liquid·Phase NMR (Academic, New York 1988) C. Corso: Physics of NMR Spectroscopy in Diology aNi MediciM, Pmc. lnl. School of Physics, cd. by B. Maraviglia (North-Holland, Amslerdam 1988) R.R. Ernst, G. Bodenhausen, A. Wokaun: PrillCiples of Nlll:lear MagflCtic Resonance in OM aNi Tl\AO DimellSiollS (Clarendon, Oxford 1981) R. Fl'eoeman: A JlaNlbo<M o{Nlll:kar Maglltlic Resanance (Wiley, New York 1987) E. FUkushima, S.B.W. Roeder: ExptrimerllalPuiseNMR_ANuulJlId BoluApproadl (AddisonWesley, Reading, MA 1981) B.C. Gerstein, C.R. Dybowski: TrlUlS~'" TecMiquu ill NMR o{Solids: All 11l1rothu:tioll 10 TIttory fJIIIi Practice (Academic, Orlando, FL 1985) U. l-laeberlen; Jligh Ruolulu:m NMR ill Solids: Sefet:li~ Avuagu., (Academic, New York 1976) P. Msnsfield, P.G. Morris: NMR lmagilll ill Biomt!dicine, ed. by J.S. Waugh (Academic, New York 1982) M. Mehring: PrillCiplu 0{ J1igh ResolUlion NMR ill Solids, 2nd ed. (Springer, Berlin 1983) M. Munowitz: CoherenceandNMR (Wiley, New York 1988) K. Wlithrich: NMR of ProteillS aNi Nucteie Acids (Wiley, New York 1986) R.N. Zare: Angular Momentum: UNierstaNlil18 Spatial Aspects ill Chemistry aNi Physics (Wiley, New York 1988)
I. Lowe: Free induction decays of rolating solids. Phys. Rev; LeU. 1, 285 (1959) E.R. Andrew, A. Bradbury, R.G. Eades: Nuclear magnetIc resonance spectra frolll a cryslal rotated at high speed. Nature 181, 1659 (1958) . E.R. Andrew, A. Bradbury, R.G. Eades: Remov...l of d.ipolar broadenmg of nuclear magnetic resonanc(' spectra of solids by specimen rotation. Natu~ 183, 1~2 (1~59) II. Kessmeier, R.E. Norberg: Pulsed nuclear magnetic resonance III rotat11l& sohds. Phy•. Rev. ISS, 321 (1967) . 0 300 3316 M.M. Maricq and J.S. Wa~h: NMR in rotatinl solids. J. Chern. Phys.7 ,3 (1979)
636
637
References
Chapter 1 1.1 1.2 1.3
C.J. Gorter, L.J.F. Broer: Physica 9, 591 (19<12) E.M. Pur«ll, II.C. Torrey, R.V. Pound: Phys. Rev. 69, 37 (1946) F. BI()(.h, W.W. nansen, M. Pac:brd , Ph)',. Rev. 69, 127 (1946)
Chapter 2 2.1 2.2 2.3 2.4
2.5 2.6 2.7 2.8 2.9 2.10 2.11 2.12 2.13 2.14 2.15 2.16 2.17 2.18 2.19 2.20 2.21 2.22
J.l1. Van Vleck: Phys. Rev. 74, 1I6S (1948) M. Tinkham : GfOIq) ThQ)ry fJIIIl QUllIIllUll Mechan ia (McGraw -Hili, New York 1964) H. Rauch, A. 7.eitinger, G. Bl'durek , A. \Vilfing, W. Bauspie$ ll, V. Boose: Ph)'s. Lett. 54A, 425 (1975) S.A. Werner, n. Colella, A.W. Overhlluser, C.F. Eagen: Phys. Rev. LeU. 35, 1053
(1975) II.J. Bernstein: Phys. Rev. Lell. 18, 1102 (1967) Y. Aharano v, L. Susskind : Phy,. Rev. ISS, 1231 (1967) M.E. Stoll, A.J. Vegll, R.W. Vaughan : Ph)',. Rev. AI6, 1521 (1977) M.E. Stoll, E.K. Wolff, M. Mehring: Ph)'s. Rev. A 17, 1561 (1978) E.K. Wolff, M. Mehring : Phys. Lett. 70A, 125 (1979) E.L. lIahn: Phys. Rev. 80, 580 (1950) II.Y. Carr: Current Comme nts 20,24 (1983) II.Y. Carr and E.M. Pun:ell: Phys. Rev. 94, 630 (1954) I. Solomon : Phys. Rev. 110, 61 (1958) J. Spobs: Thesis. University of Illinois (1957) (unpubli sll«!.) J.J. SpokM, C.P. Slicht.cr: Phys. Rev. 113, 1462 (1959) D.F. Holcomb , R.E. Norberg : PhYlI. Rev. 98,1074 (1955) Z. Wang: Private commun ication C.J. Gorter: ParanulgM/icRtlaxatiorl (Elsevie r, New York 1947) p. 127 R. Kubo, K. Tomita: J. PhYll. Soc. Japan 9, 888 (1954) E. W. Hobson: TM TMoryo f FlUlClioru ofa Rwl Variable aIId 1M Theory of FOfUids Strits (Cambri dge UnivcrlIity Press, Cambrid ge 1926) p. 353 IT. P.W. Anderso n: J. Phyll. Soc. Jllpan 9, 316 (1954) 11.5. Gutowsk y, G.E. Pake: J. Chern. PhYll. 16, 1164 (1948), ibid. 18, 162 (1950) for example
Chapte r 3 3.1
E.n. Andrew , n.G. Elides: Proc. n. Soc., London A1.18, 537 (1953)
Chaple r 4 -4.1 4.2
N.r. namsey : Phys. Rev. 78, 699 (1950); Phys. Rev. 86, 213 (1952) W. Lamb: PhYll. Rev. 60, 817 (1941)
639
...
<.3 <.5
<.6
... <.•
".10 4.11 4.12
U3 ".14 4.15 4.16 4.17 4.l8 4.19 4.20 1.21 4.22 4.23
J.A. Pople: Proc. R.. Soc., London A1J9, 5-41, 550 (1957) N.r. Ramsey: Phys. Rev. 86, 243 (195'2) D. Pines. Solid Stale Physics, Vol. I, ed. by F. Seib, D. Turnbull (Academic, New York 1955) L. Pauling, E.B. Wilson, Jr.: lnlroductionlO Quan/um Mechanics (McGraw-Hili, New Vork 1935) p. 232 ILT. Schumacher, C.P. Slichter: Ph)'•. Rev, lOt, 58 (1956); B.R. Whiting, N.S. Van der Ven, R.T. Schumacher: Phys. Rev. B18, 5413 (1978) Ch. Ryter: Ph)'.. Rev. LeU. 5, 10 (1960) W. Kohn: Ph)'•. Rev. 96, 590 (l9r.4) T. Kjeldaas, Jr., W.Kohn: Ph)'•. Rev. 101,66 (1956) S.II. Vosko, J.P. Pcradew, A.l1. MacDonald: Ph)'•. Rev. LeU. 35,1725 (1975) A.C. Gossard, A.M. Portis: Ph)'s. Rev. Lett. 3, 1M (1959); J. Appl. Ph)'•. Suppl. 31,2055 (1960) W. Marshall: Ph)'•. Rev. 110, 1280 (1958) R. Walson, A. Freeman: Ph)',. Rev. 113, 2021 (1961) N. Bloembergen, T.J. Rowland; Ad.. ~letl.1l. 1, 731 (1953) E.L. Ibhn, D.E. MuW('lI: Phys. Rev. 88, 1070 (1952) H.S. Gutowsky, D.W. "'lcCaIl, C.P. Slkhler: J. Chern. Phys. 11, 219 (1953) N.F. Ramsey, E.M. Purcell; Phys. Re.... 85, 143-144 (1952) (Idter) N. Bloembergcn, T.J. Rowland: Phys. Re.... 97, 167~1698 (1955) M.A. Ruderman, C. KiHei; Phys. Rev. 96, 99 (19M) F. Frohlich, F.R.N. Nabarro: Proc. R. Soc. London AI7S, 382 (1940) K. Yoshida: Phys. Rev. 106,893 (1957) J.II. Van Vleck: Rev. Phys. 34, 681 (1962)
Chapter 5 5.1 5.2 5.3 5.'l 5.5 5.6 5.7 5.8 5.9 5.10 5.11 5.12 5.13 5.14 5.15 5.16
C.J. Gorter; ParQmllfic RtfcuatwII (Elsevier, New York 1947) L.C. Hebel, C.P. Slichter: Phys. Rev. 113, 1504 (1959) J. Korrillga; Physica 16, 601 (1950) D. Pines: Solid Stall Physics, Vol. I, cd. by F. Seit:!:, D. Thmbull (Academie, New York 1955) R.C. Tolman: T~ Principles of Skllistical Mtchania (Odord University Press, New York 1946) I.J. Lowe, R.E. Norberg: Phys. Rev. 107, 46 (1957) C.R. Bruce: Phyt. Rev. 107,43 (1951) R.R. Ernst, W.A. Anderson; Rev. Sci. Instrum. 37, 93 (1966) R.R. Ernst, G. BodenhauS
Chapter 6 6.1 6.2 6.3 6.4 6.5
II.B.G. CllSimir, F.K. duPri: Physica 5, 507 (1938) J.II. Van Vleck: Phys. Rev. 57, 426 (1940) J.II. Van Vleck: J. Chern. Phys. 5, 3'20 (1937) L Waller: Z. Phys. 79, 370 (1932) A.G. Redfield: Phys. Rev.1l8, 1787 (1955)
6.' 6.7 6.8
6.'
6.10 6.11 6.12 6.13 6.14 6.15 6.16 6.17 6.18 6.19 6.20 6.21 6.22 6.23 6.24 6.25
C.P. S/ichter, W.C. Holton; Phys. Rev. 122, 1701 (1961) E.I!. Thrner, A.M. Sachs, E.M. Purcell: Phys. Rev. 76, 465 (A) (1949) R.V. Pound: Phyt. Rev. 81,156 (1951) N.F. Ramsey, R.V. Pound; Phys. Rev. 81, 278 (1951) . R.V. Pound, E.M. Purcell: Phys. Rev. 81, 279 (1951) J.II. Vsn Vleck: Nuovo Ciment.o Suppl. 6, $erie X, 1081 (1957) L.C. Hebel, C.P. Slichler: Phys. Rev. 113, 1504 (1959) A.G. Redf)Cld: Phys. Rev. Leu. 3, 85 (1958) A.G. Redfield, A.G. Anderson: Phys. Rev. 116, 583 (1959) R.W. Morse, II.V. 8ohm: Phya. Rev. 108, 1094 (1957) L. Cooper: Phys. Today 26, (31 July 1973) A.G. Anderson: Phys. Rev. liS, 863 (1959) A. Abragam, W.G. Proctor: Phys. Rev. 106, 160 (1957) A.G. Redfield; Phys. Rev. 98, 787 (1955) C.P. Slichter, W.C. lIo/lon: Phys. Rev. 122, 1701 (1961) B.N. Provotorov: Soviet Phys. - JETP 14,1126 (1962) R.T. Schumacher: Phys. Rev. 112, 837 (1958) GP. Slichter, D. Ailion: Phyt. Rev. I3S, A 1099 (1964) D. Ailion, C.P. Slichtcr: Phya. Rev. 137, A235 (1965) D.C. Look, I.J. Lowe: J. Chem. Phya. 44, 2995 (1966)
Chapter 7 7.1 7.2 7.3
7.< 7.5
7.'
7.7 7.8 7.' 7.10 7.11 7.12 7.13 7.14 7.15 7.16 7.17 7.18 7.19 7.20 7.21 7of2 7.23 7.24 7.25 7.26 7.27 7.28 7.29 7.30 7.31 7.32
R.V. Pound: Phys. Rev. 79, 685 (1950) T.R. Carver, C.P. Slichter: Phys. Rev. 92, 212 (1953); ibid. 101, 975 (1956) A.W. OV~hauser: Phya ~ev. 91, 476 (1953), ibid. 92, 411 (1953) C.D. Jeffnes (~); ~fIilJ1UC.NIIC~Or~~at~ (Inlerscience, New York 1003) D.F. lIolcomb. ThOlIS, Umvennty of IlhnOls (1954) unpublished F. Bloch: Phya. Rev. 102, 104 (1956) 1. Solomon: Phys. Rev. 99, 559 (1955) C.D. Jclfries: Phys. Re~. 106, 164 (1957); ibid. 117, 1056 (1960) A. Abragam, J. Combnsson, I. Solomon: C.R. Acad. Sci. 247, 2237 (1958) E. Erb, J.L. Montchane, J. Uebersfeld: C.R. Aead. Sei. 246, 2237 (1958) G. Fel)Cr: Phys. Rev. 105, 1122 (1957) II. Seidel, II.C. Wolf: In PJrysiai¥CoIorCtnlus, ed. by W. Beall Fowler (Ae..demic New York 1968) , J.P. Gordon, H.J. Zeiger. C.II. Townes; Phya. Rev. 9S, 2821 (1954) Nobel Ltctwres-Physu:s 1963-1970 (Elsevier, Amsterdam 1972) N.G. Bll.SOv, A.M. Prokhorov; Sov. Phys. - JETP 27,431 (1954) N. Bloclllbergen: Phys. Rev. 104, 324 (1956) A. Abrsgam, W.G. Proctor: Phys. Rev. 106, 160 (1957) N. Bloembergen, S. Shapiro, P.S. Pershan, J.O. Artman: Phys. Rev. 114, 445 (1959) p.s. Pershan: Phys. Rev. 117, 109 (1960) A.G. Anderson: Phys. Rev. IlS, 863 (1959) A.G. RcdflC.ld: Phyt. Rev. 130,589 (1963) N.C. Fernehus: Proe. of the XIV Colloque Ampere (1966) p. 497 R.E. Slusher, E.L. nahn: Phys. Rev. 166, 332 (1968) M. Minier: Phys. Rev. 182, 437 (1969) ~~37eitekaml)' A. Bielecki, D. Zn, K. Zilm, A. Pines: Phya. Rev. Lelt. SO, 1807 D.T. Edmonds: Phys. Rep. 211, 233 (1977) N. Bloembergen, P. Sorokin: Phys. Rev. 100,865 (1958) S.R. lIarimann, E.L. lIahn: Phys. Rev. 128, 2042 (1962) ".M. Lurie, GP. Slichler: Phys. Rev. 133, AlI08 (1964) c.P. Slichler, W.C. lIolton: Phys. Rev. 122, 1701 (1961) P.R. Spencer, N.D. Schmid, C.P. Slichter: Phys. Rev. 81, 2989 (1970) D.V. Lang, P.R. Morl\fl: Phys. Rev. III, 53 (1970)
640 641
7.33 7.34 7.35 7.36 7.37 7.38 7.39 7.40 7.41 7.42 1.43 7.44 7.'l5 7.46 7.47 7.48 1.49 7.50 7.51 7.52 7.53 7.54 7.55 7.56 7.57 7.58 7.59 7.60 7.61 7.62 1.63 7.64 7.65 7.66 7.67 7.68 7.69 7.70 7.71 7.72 7.73 7.74 7.75 7.76 7.77 7.78 7.79 7.80 7.81 7.82 7.83 7.84
642
A. Pinell, M.G. Gibby, J.S. Waugh: J. Chem. Phys. 56, 1776 (1972); ibid. 59, 569 ( 1973) V. Royden: Phys. Rev. 96, 543 (1954) A.L. Bloom, J.N. Shoolcry: Phys. Rev. 97,1261 (1955) F. Bloch: Phys. Rev. 93, 944 (1954) R. Freeman, S.P. Kernpsel1, M.H. Levilt: J. Magn. Reson. 35, 447 (1979) W.A. Anderson, F.A. Nelson: J. Chern. Phys.39, 183 (1963) R. Freeman, W.A. Anderson: J. Chern. Phys. 42, 1199 (1965) R. Ernst: J. Chem. Phys. 45, 3845 (1966) J.B. Grut.ner, R.E. Santini: J. Magn. Rcson. 19, 173 (1975) V.J. Basus, P.O. Ellis, II.D.W. lIi11, J.S.WlIugh: J. Magn. Rcsoll. 35,19 (1979) M.II. l.evitt, R. forcernall: J. Magll. Rcsoll. 33, 413 (1979) M.II. Levitt: Prog. Nud. Magn. Reson. Spcctrosc. 18, 16 (1986) A.J. Shalea, J. Keeler: Prog. Nuc!. Magn. Reson. Sped~. 1.9, 47 (1987) R. Freeman: Bull. Magn. Reson. 8, 120 (1986) D.E. Kaplan, E.L. lIahn: J. Phys. Radium 19,821 (1958) M. Emahwiller, E.L. lIahn, D. Kaplan: Phys. Rev. liS, 414 (1960) J .B. Boyce: Thesis, University of Illinois (1972) D.V. Lang, J.B. Boyce, D.C. 1,0, C.P. Slichter: Phys. Rev. Lett. 29, 716 (1972) D.V. Lang, D.C. 1,0, J.B. Boyce, C.P. Slichter: Phys. Rev. 09, 3071 (1914) C.D. Makowb, C.P. Stichter, J.II. Sinfelt: Phys. Rev. Lett. 49, 379 (1982); Phya. Rev. 831, 5663 (19M) P.K. Wang: Thesis, University oflllillOis (1984) unpublished P.K. Wang, C.P. Slidlter, J.II. Sinfelt: Phys. Rev. Lett. 53, 82 (1984) J. Jeener: Lecturu, Ampere Intemationa.l Summer School, Basko Po~u, YugOll1avia (1971) unpublished R. Freeman, G.A. Morris: Bull. Magn. Rcson. 1, 5 (1979) W.P. Aue, E. Bartholdi, R.R. Ernst: J. Chem. Phys.~, 2229 (1976) A. Bax: Bull. Magn. Reson. 78,167 (1985) D.L. Thmer: Pros. Nud. Magn. Reson. Spcctrosc. 17, 281 (1985) L. Muller, A. Kumar, R.R. Ernst: J. Chern. Phys. 63, 5490 (1915) L. Muller, A. Kumar, R.R. Erllllt: J. Magn. Res<m. 25, 383 (1977) D.T. Pegs, M.R. Bendall, D.O. Doddrell: J. Magn. Re8on. 44, 238 (1981); ibid. 45, 8 (1981) G. Feher: Phya. Rev. 103, 500 (1956) G. Feher, E.A. Gere: Phya. Rev. 103, 501 (1956) E.B. Baker: J. Chem. Phyll. 37, 911 (1962) K.G.R. Pllchler, P.L. Wessels: J. Magn. Reson. 12,337 (1973) A.A. Maudsley, R.R. Ernst: Chem. Phys. Lett. SO, 368 (1977) G. Bodenhausell, R. Freeman: J. Magn. Rcson. 28, 463 (1977) A.A. fo,iaudsley, I,. Muller, R.R. Ernst: J. Magn. Rcson. 28, 463 (1977) G.A. Morris, 11.. Freeman: J. Am. Chern. Soc. 101, 760 (1979) O.W. Sorensen, G.W. Eich, M.JI. Levitt, G. Bodenhauscn, R.R. Erno,st: Prog. Nuel. Magn. Reson. Spcetrose. 16, 163 (1983) F.J.M. Van den Ven, C.W. IIilbcrs: J. Magn. Reson. 54, 512 (1983) P.K. Wang, C.P. Slicht.er: Bull. Magn. Reson. 8, 3 (1986) E.I,. lIahn, D.E. Maxwell: I>h)'s. Hev. 88, 1070 (1952) 11.5. Gutowsky, D.W. McCall, C.P. Sliehler: J. Chem. PhYll. 21, 279 (1953) K. Wuthrich: NMR of Proteins and Nucleic Acids (Wiley-lntcrscience, New York 1986) P. Mansfield: NMR J!IUI8ing in Biomedicine (Academic, New York 1982) R.R. Ernst, G. Bodenhausen, A. Wokaun: Principles of Nuclear r.hgnetic l1.esonance in One and Two Dimensions (Clarendon, Oxford 1987) P.C. Lalilerbur: Nature 242, 190 (1973) P. Mansfield, P.K. Crannell: J. Ph)'s. C6, L422 (1973) A. Kumar, D. Welti, R.R. Ernst: J. Magn. Reson. 18, 69 (1975) P. Mansfield: J. Phys. C 10, L55 (1971) K.K. King, P.R. Monln: Med. Phys. II, I (1984) T.R. Brown, B.M. Kincaid, K. Ugurbil: Proc. Natl. Acad. Sci. USA 79, 3523 (1982)
7.85 7.86
0.8. Twieg: Moo. Phys. 10 610 (1983) S.J. Ljunggren: J. Mllgn. R~n. 54, 38 (1983)
Chapter 8 8.1 8.2
8.3 8.' 8.' 8.8 8.7 U 8.9
£.L. Hahn: Phya. Rev. 80, 580 (1950) H.Y. CarT, E.M. Pun:ell: Phys. Rev. 94, 630 (1954) S. Meiboom, D. Gill: Rev. Sci. Instrum. 29, 6881 (1958) E.D. Ostroff, J.s. Waugh: Phys. Rev. Lett 16, ioo7 (1966) P. Mansfield, D. Ware: Phys. Rev. Leu. 21,133 (1966) t.M. Stacey, R.W. Vaughan, D.O. Ellernan: Phya. Rev. Lett. 26, 1153 (1971) J.G. Powles, P. Mansfield: PhYI. Lett. 2, 58 (1962) I. Solom~n: Phya. Rev. 110, 61 (1958) ~9:'(?9a;6)' K.R. Jeffrey, M. Bloom, M.I. Valie, T.P. lIiMB: Chem. Phya Lett. 42,
8.10 P.K. Wang: Thesis, University of Illinois (1984) unpublished 8.11 P.K. Wang, C.P. Sliehter: Bull. Magn. Reson. I, 3 (1986) 8.12 J. Jcener, P. 8roenert: Phya. Rev. 157, 232 (1967) 8.13 n96t.;)' W.1. GoJdburl: Phys. Rev. Leu. H, 2SS (1963); Phya Rev. 140, AIUl D. Ba.rnaal, I.J. Lowe: Phys. Rev. Leu. 11, 258 (1963) W.-K. Rhim, H. Kcssemeier: Phys. Rev. OJ, 3655 (1971) H. Kcsse~ier, W.:K. Rhim: Phys. Rev. B5, 761 (1972) R W.-K. Rhlm, A. Pllles, J.S. Waugh: Phya. Rev. Lett 2S 218 (1970)· Ph BJ,684 (1971) " , ya. ev. 8.18 I. Solomon: Phya. Rev. Lelt. 2, 301 (1959) 8.19 K. Takegoshi, C.A. McDo_II: Chern. PhYI. Lett. 116, 100 (1985) 8.20 U.S. GUlowsky, G.E. Pake: J. Chem. Phy•. 16, 1164 (1948)' ibid. 18 162 (1950) 8.21 I. Lowe: Phys. Rev. Lett. 2, 285 (1959) " 8.22 g:i9)ndrew, A. Bradbury, R.G. Eades: Nature 182, 1659 (1958); ibid. 183, 1802 8.14 8.15 8.16 8.17
8.23
:ri~~:~~:I~~)farMomell/lUnillQlUlflIumMtduvUcs (Princeton University Press,
II. Kcssemeier, R.E. Norberg: Phys. Rev. 155, 321 (1967) M.M. Ma~cq, J.S. Waugh: J. Chem. Phys. 70, 3300 (1919) ~.M. Po~tlll: Phys. Rev. 91,1971 (1953) .0. Stejskal, J. Schaefer, R.A. McKay: J. Magn. Reson. 25, 569 (1917) ~ Herdeld, A. RoufOSS(!, R.A. lIaberkorn, R.G. Griffin, M.J. Glimcher' Phi los. al18. R. Soc. London, Series B 2&9, 459 (1980) . 8.29 J. He~~eld, A.E. Berger: J. Chern. PhYI. 7J, 6021 (1980) 8.39 \V.T. Dlx?n: J. Chem. Phya. 77, 1800 (1982) 8.31 (~:8·.R) ',Ielg,h, E.T. Olejniuak, S. Vega, R.G. Griffin: J. Am. Chern. Soc. 106 8302 ; : fag? Reson. 72, 238 (1987) , 8.32 M. Mehrmg: /flgh ResoluJion NMR S~c//"O$co'"' in Solids NMR· B . p. . I Pro V I II ed b ' Y J , . l\Sle rlllClp es lin d 197:)' o. , . y P. DIehl, E. Fluck, R. Kosfcld (Springer, Berlin, Heidelberg 8.24 8.25 8.26 8.27 8.28
8.34 8.35 8.36 8.37
U. Haeberle.n: IJigh ~esolu/ion NMR in Solids: Selec/ive Averaging; Supplement No. lo Advancts If! Magnellc Resonance (Academic, New Yorl.: 1976) J.S. Waugh, L.M. Huber, U. HaeberJen: Phys. Rev. Lett. 20,180 (1968) U. Ha.eberlen, J.S. Waugh: Phys. Rev. 175,453 (1968) P. Mansfield: J. Phya. C4, 1444 (1971) ri9~)"sJ1eld, M.J. Orchard, D.C. Stalker, K.II.B. Richards: Phys. Rev. B7, 90
8.38 8.39
~9~'(~~~)'
8.40
L. Van Gerven (ed.): Nl«.'.ftar MagntlU; RUOfICIlICt in Solids (Plenum, New York 1977)
8.33
W.K. Rh~m, D.O. Ellernan, R.W. Vaughan: J. Chern. Phys. 59 3740 (1973) D.O. ElJeman, L.B. Seh~iber, R.W. Vaughan: Chern. Phys. 60,
i.
643
8.41 8.42
P.R. Moran: J. Phys. Chern. Solids 30,297 (1969) W.S. Warren, S. Sinton, D.P. Weitckarnp, A. Pines: Phy•. Rev. Lett. 4], 1791 (1979)
Appendix F.l
F.7 F.8 F.9
H.S. Gulowsk)', C. Holm: J. Chern. Phy•. 25, 1228 (1956) H.S. Gulowsky, D.W. McCall, C.P. Slichter: J. Chern. Phy•. ll, 279 (1953) .B.L. Hahn, D.E. Maxwell: Phy•. Rev. 88, 1070 (UI52) J.lI. Van Vied:: Ned. Tijdschr. Natuurkd. 27,1 (1961) H.M. McConnell: J. Chern. Phy•. 28, 430 (1958) D.H. Archer: Thesis, lIarvard University (1953) P.W. Anderson: J. Phy.. Soc. Jpn. 9, 316 (1954) H.S. Gulowsky, A. Saib.: J. Chern. Phys. 21, 1588 (19S3) J.T. Arnold: Phys. Rev.IOl, 136 (1956)
G.I G.2
H.C. TOrTey: Phys. Rev. 104, 563 (1956) H.Y. Carr, E.M. Pu«:ell: Phy•. Rev. 94, 630 (J954)
I.l 1.2
1.6
N. Bloembe:r'!en, T.J. Rowland: Acta Metall.l, 731 (1953) 1.5. Gradshte)'n, I.M. R)'llhik: Tabu 0/ in/llrols, Sums, &riu, and Products, 4lh ed. (Academic, New York 1980) pp. 177, 904 B.T. Feld and W.E. Lamb: Phys. Rev. 67, 15 (1945) N. Bloembe:rgtln: In Dt!ICts in errS/alUM Solids (Phy.ical Society, London 1954) pp. 1-32 D. Gerstein, C. Dybowski: Tnlflsien/ TechniqulS in NMR oj Solids (Academic, New York 1985) G.E. Pake: J. Chern. Phy•. 16,327 (1948)
J.l
F.J. Dyson: Phys. Rev. 75, 486 (1949)
K.I K.2 K.3 K.4 K.5
U. Haebe:rlen, J.S. Waugh: Phy•. Rev. 125,453 (1968) R.M. Wilcox: J. Math. Phy•. 8, 962 (1967) W. Magnw;: Commun. Pure Appl. Malh. 7, 649 (1954) P. Mansfield: J. Phya. e4, 1444 (1911) D.P. Burum, M. Linder, R.R. Ernst: J. Ma&". Raon. 44, 173 (1981)
F.2
Chapter 9
F.3 F.4 F.5 F.6
G. Bodenhausoen: Pr<'>fl;. Nucl. MaSn. Reson. Sp«trosc. 14, 137 (1981) II. lIataoab, T. Terao, T. lIa.shi: J. Phys. Soc. Japan 39, 835 (1975) II. Hatanab, T. lIashi: J. Ph)'•. Soc. Japan 39, 1139 (1975) W.P. Aue, E. Biutholdi, R.R. Ernst: J. Chern. Ph)'•. 601, 2229 (1976) D. Weitebmp: Adv. Magn. R.eson 11, 111 (1983) S. Emid: Bull Magn. Reson. 4, 99 (1982) M. Munowib, A. Pines: Science 233, 525 (1986); Adv. Chern. Phy•. LXVI, I (1987) R.R. Ernst, G. Bodenhausen, A. Wobun: Principiu ojNlU:lnuMug~/U:Ruonanceill One and Two DilMnsiotu (Clarendon, Oxford 1987) 9.9 M.E. Sloll, A.J. Vega, R.W. Vaughan: Phys. Rev. A16, 1521 (1977) 9.10 M.E. Sloll, E.K. Wolff, M. Mehring: Phy!!. Rev. A17, 1561 (1978) 9.11 E.K. Wolff, M. Mehring: Phy•. Lett. 70A, 125 (1979) 9.12 M. Mehring, P. HOfer, A. Grupp: Phy!!. Rev. A3l, 3523 (1988) 9.13 P.K. Wang, C.P. Slichter, J.II. Sinrelt: Phys. Rev. Lett. 53, 82 (1984) 9.11 G.E. Pake: J. Chern. Phy•. 16, 327 (19<18) 9.15 J. Baum, M. Munowitll, A.N. Garroway, and A. Pines: J. Chelll. Phy•. 83, 2015 (1985) 9.16 E. Menba.::her: Quan/um MecJw.nics, 2nd cd. (Wiley, New York 1970) p. 167 9.17 A. Wokaun, R.n. Ernst: Chcm. Phy•. Lelt. 51, 407 (1977) 9.18 W.S. Warren, S. Sinlon, D.P. Weitelcamp, A. Pines: Phy•. Rev. Lett. 43, 1791 (1979) 9.19 W.S. Warren, D.P. Weitcbmp, A. Pines: J. Chern. PhYll. 73, 2084 (1980) 9.20 Y.-S. Yen, A. Pines: J. Chern. Phy•. 73, 3579 (1983) 9.1 9.2 9.3 9.4 9.5 9.6 9.7 9.8
1.3 1.4 1.5
Chapter 10 10.1 10.2 10.3 10.4 10.5
J.H. Smith, E.M. Purcell, N.F. Ramsey: PhYll. Rev. 108, 120 (1957) L.R. Walker, G.K. Wertheim, V. Jaccarino: Phy•. Rev. Lett 6, 98 (1961) M.E. R.o5e: Efmwllaryl'Moryoj ....nguuuMf)tMn1IInc (Wile:y, New York 1957) E. Ambler, J.C. Eisenstein, J.P. Schooley: J. Math. PhYll. J, 118, 760 (1962) R.M. Stemheimer: Phys. Rev. 84, 244 (1951); ibid. 86, 316 (1952)j ibid. 95, 736 (1954)
Chapter 11 11.1 11.2 11.3 IIA 11.5 11.6
644
A.R. Edmonds: AflgulorMomefl/umifl Quof1lumMechaflit;s (PrinCtllOn Univtlrsily Press, Princeton, NJ 1957) J.S. Waugh, C.P. Slichter: Phy•. Rev. B37, 4337 (1988) I,.G. Rowan, E.L. Hahn, W.B. Mims: Phys. Rev. 137, 61 (1965) W.B. Mim.s: Phys. Rev. B5, 2409 (1972) T.G. ClISlncr, W. Kin~i8: J. Phy•. Chem. Solids. J, 178 (1957) CoJ. Delbceq, B. Smaller, 1'.11. Vusler: Ph)'s. Rev. 111, 1235 (1958)
645
Author Index
Abragam, A. 231,264,271 Aharanov, Y. 32 Ailion, D. 245 Ambler, E. 494 Anderson, A.G. 230,274 Aooerson, P.W. 63 Anderson, W.A. 179,183,
31' Andrew, E.R. 80,82,85, 215,392,399,405 Ansermet, J.-P. 301 Archer, D.H. 593 Arnold, J.T. 597 Artman, J.O. 272 Aue, W.P. 319,324,325, 433 Badurek, B. 32 Baker, E.B. 331,351 Bardeen, J. 231
Boyce, J.B. 312 Bradbury, A. 392,399, 405
Broekaert, P. 380,381, 383,384 Broer, L.J.F'. 9 Brown, T.R. 364 Bruce, C.R. 179 BUfum, D.P. 628 Caff, H.Y. 10,11, 367-369,392,404,601 Cafver,T.R. 219,250,257 Casimir, U.B.G. 219 Castner, T.G. 533,534 Cohen, M.H. 485 Colella, R. 32 Combri880n, J. 264 Cooper, L.N. 231
Bunaal, D. 387
Das, T .P.
Bartholdi, E. 319,3z.t, 325,433 Basov, N.G. 269,270 BMUS,VJ. 310 Baum, J. 448 Ba.uspiess, W. 32 Bax, A. 319,325 Bendall, M.R. 331 Bernsle;n, H.J. 32 Biel«.ki, A. 27<1 Bloeh, F. 9,35,199,258, 297,309 Bloembergen, N. 127,133,
Davis, J.II. 312 DeIBecq, C.J. 533 Dixon, W.T. 406 Doddrell, D.O. 331 du Pre, F.K. 219 Dun.nd, D.J. 301 Oybowsld, C. 616 Dyson, F. 617,618
14 1, 143, 199, 206, 231,
269,270,275,276,281, 284,289,290,292,291, 606,615 Bloom, A.L. 297,307, 309,310 Bloom, M. 372 Bodenhausen, G. 179, 331,340,342,345,350, 354,357,433,434 Bohm, D. 115,124 Bohm, H.V. 231 Boose, V. 32
485
Eades, R.G. 80,82,85, 215,392,405 Edmonds, A.R. 522 Edmonds,D.T. 275 Eich, G.\\'. 345,350,354 Eisenstein, J.G. 494 E11eman, D.O. 367,406, 422
Ellis, P.D.
Emid, S.
310
434
Emshwiller, M. 311 264 Ernst, R.E. 179,183,310. 319,322,323-325,331, 335,338,345,350.354, 357,362,363,433.434. 463,464,469, 4i3, 628
Erb, E.
Feher, G. 266-268,331, 332,335,337 Feid, B.T. 615 Fernelius. N. 274 ....r eeman. A. 126 Freeman, R. 307,310,311, 319,325,331,340,342 "'rohlieh, F. 133,143 Garroway, A.N. 448 Gefslein, D. 616 Gibby, M.G. 247,293 Gill, D. 367,369,371,422 Glimcher, M.J. 405 Goldburg, W.1. 384,386, 388,389 Gordon, J.P. 269 Gorter. C.J. 9,52,151,
21. GO!l$l'lrd, A.C. 126 Cradshleyn, 1$. 612 Granell, P.K. 358,361 Griffin, R.C. 405,406 Grupp, A. 440 Grutzner, J.B. 310 Gutowsky, M.S. 63,80, 127,132,141,351,392, 592,593,596 Haberkorn, R.A. 405 Ih.eberlen, U. 421,422, 423,624,627 Ibhn, E.L. 39,40,43,127, 132,247,274-277.283. 287,289,293,311,351, 367,485,524,593,601 Hansen, W.W. 9 Hartmann, S. 247, 275-277.283,289,293 lIashi, T. 433 lIatanaka. H. 433 liebel, L.C. 155,230 lIen:rekl. J. 105 IIiggs, T.P. 372 HUbers, c.w. 345,354, 356
647
Hill, H.D.W. 310 Hobson, E.w. 56 HOfer, P. 440 Holcomb,D.F. 42,44, 215,258 Holm, C. 592,593 Holton, W.C. 220,235 Huber, L.M. 422,423 Jaccarino, V. 487 Jeener, J. 319,331,335, 351,380,381,383,384 Jeffre~ K.R. 372 Jeffries, C.D. 264 Kii.nzig, W. 533,534 Kaplan, D.E. 311 Keeler, J. 311 Kempsell,S.P. 307,310 Kessemeier, II. 389,391, 398,399 Kincaid, B.M. 36-1 King, K.K. 364 Kittel, C. 133,143 Kjeldus, T. Jr. 123,124 Knight, W. 113 Kohn, W. 123,124 Korringa, J. 156,157 Kubo, R. 55 Kumar, A. 322,323,362, 363 Lamb, W. 103,615 Lang, D.V. 287 Lau~erbur, P.C. 357,358, 361 Lee, M. 384,386,388,389 Levitt, M.II. 307,310, 311,345,350,354 Linder, M. 628 Ljunggren,S.J. 364 Look, D.C. 245 Lowe,I.J. 179,183,245, 387,392,398 Lurie, F. 275,281,283, 284,286,287,289,293 Magnus, W. 475,478, 624,625 Makowka, C.D. 312,313 Mansfield, P. 357,358, 361, 362, 36~367, 371, 372,406,628 Maricq, M.M. 399 Marshall, W. 126 Maudsley, A.A. 331,335, 338 Maxwell, D.E. 127,132, 351,593
648
McCall,D.W. 127,132, 141,351,593 McConnell, H.M. 593 McDowell, C.A. 392 McK ..y, R.A. 4(H,406 Mehring, M. 33,440 Meiboom, S. 367,369, 371,422 Merzbacher, E. 459 Mims, W.B. 524 Minier, M. 274 Monteh ..ne, J.L. 264 Moran, P.R. 287,364,422 Morris, G.A. 319,325, 331 Morse, R.W. 231 Muller, L. 322,323,331 Munowilt, M. 434,448, 463,471 Nabarro, F.R.N. 133,1<13 Nelson, F.A. 310 Norberg, R.E. 42,179, 183,215,398,399 Olejniczak, E.T. 406 Orcharo, M.J. 406 Ostroff, E.D. 367 OverhaU&er, A.W. 32, 247,249,250,254,256, 257,293-295 Puhler, K.G.R. 331,338, 351 Pa<:kard, M. 9 Pake, G.E. 63,80,392, 448,616 Pauling, L. 117 Pegs, D.T. 331 Pennington, C.II. 301 Persh,,", P. 272-274 Pines, A. 247,293,389, 391,392,434,448,463, 471,475,477,479,480,
'83
Pines, D. 115,124,157 Pople, J.A. 105 Por~is, A.M. 126,399 Pound, R.V. 9,199,206, 229-231,247-250,295 Powles, J.G. 371,372 Proctor, W.G. 231,271 Prokhorov, A.M. 269,270 Provolorov, B.N. 239,240 Purcell, E.M. 9,40,401, 132,199,206,230,231, 367-369,392,404,486, 661
Raleigh, D.P. 406 Ramsey, N.F. 92,107, 132,229,486,511 Rauch, H. 32 Redfield, A.G. 34,199, 200,201,204-206, 219-221,230,232, 234-236,241,242,244, 274,276,279,281,289 Reif, F. 485 Rhim, W.-K. 389,391, 392,406,422 Richards, K.II.B. 406 Rose, M.E. 490,492 Roufosse, A. <105 Rowall, L.G. 524 Rowland, T.J. 127,133, 141,143,606,615 Royden, V. 297,307, 309 Ruderm..n, M. 133,143 Ryter, Ch. 122-121 Ryzhik,I.M. 612 Sachs, A.G. 229 Saika, A. 596 Santini, R.E. 310 Sehaefer, J. 404,406 Schmid, N.D. 287 Sehoolery, J.N. 297,307, 309,310 Schooley, J.F. 494 Schreiber,I,.B. 406,422 Schrieffer, J.R. 231 Schumacher, R.T. 122-124,240,241 Seidel, II. 268 Seymour, E.F.W. 215 Shaka, A.J. 311 Sh..piro, S. 272 Sin felt, J.II. 312,313,318, 444,448 Sinton, S. 428,471,479 Slichler, C.P. 42,122-124, 127,141,155,215,220, 230,235,245,249,2&0, 257,275,281,283,284, 286,287,289,293,312, 313,318,345,351,378, 378,382,444,448,524,
Sorokin, P. 275,276,281, 284,289,290,292,294 Spencer, P.R. 287 Spokas, J. 42,215 Stacey, L.M. 367,406 Stalker, D.C. 406 Slejskal, E.O. 404,406 S~ernheimer, R.M. 502 Sloll, M.E. 33,440 SU8IIkind, L. 32 Takegoshi, K. 392 Terao, T. 433 Theiss, G. 398 Tinkham, M. 32 Tolman, R...C. 157 Tomila, K. 55 Torrey, H.C. 9,598 Townes, C.H. 269 Thmer, D.L. 319,325 Thrner, E.H. 229 Twieg, D.B. 364 Uebersfeld, J. 264 Ugurbil, K. 364
Valie, M.I. 372 Van cler Yen, F.J.M. 345, 354,356 Viln Gerven, L. 421 Van Heeke, P. 421 Van Vied, J.H. 14,79, 133,142,219,221,230, 593
Vaugh..n, R.W. Vega, A.J. 33 Vega, S. 406
Weitekamp, D.P. 274, 428,434,463,471,476, 477,483 Welti, D. 362,363 Werner, SA. 32 Wertheim, G.K. 487 Wessels, P.L. 331,338,
3S1 406,422
Walker, L.R. 487 Waller, I. 219 Wang, P.-K. 318,345,378, 382,444,448 Wllng, Z. 45 Wangsnesa, R. 199 Ware, D. 367,406 Warren, W.S. 428,471, 476,477,479,483 Watson, R. 126 Waugh, J.S. 247,293,310, 367,389,391,392,399, 422,423,624,624,627
Wilcox, R.M. 624 Wilting, A. 32 Wilson, E.B., Jr. 117 Wobun, A. 179,357,434, 463,464,469,473 Wolf, H.C. 268 Wolff, E.K. 33,440 Wuthrich, K. 351 Yen, W. 398 Yen, Y.-S. 471,475,480 Yosida, K. 133,142 Yuster, P. 633 Zax, D. 274,301 Zeiger, H.J. 269 Zeilinser, A. 32 Zilm, K. 274
593
Slusher, R.E. 274,287 Smith, J.II. 486 Solomon, I. 42,258,262, 264,372,390 Sorensen,O.W. 345,350, 3&<
..9
Subject Index
I
Absorption - elementary theory 5ff - expression for power absorbed 38 - genen] ",tomie theory 59ft' - introduced I) Adiabatic changes in the mllsneLK: field 24,223ff Adiabatie dell'. .gnetintion in the lab frame 223tr - in the rotating (nune (ADRF) 23>W Adiabatie inversion 24 Alternating magnetic field, st:~ Ih Anisotropy (crystalline) - ofchemiCilI and Knight shifis 127ff - effect on powder pllUerns 60Slf Antiferromagnets - NMR in 124f Antishielding factor 502 Asymmetry parameter, defined 497 Average lJamiltonian calculated for various pulse sequences 4231f
correctiol15 to the theory of introduced .ol ... lf
623lf
Benune, structure $tOOied by second moments 80ff Bloch equatiol'l$ applied to diffusion 597lf applied to rate processes S92IT failure under some conditions 220 introduced 33 modified for relaxations in the ins~antaneous field 217 solution in low 111 3511' Bloch functions, deAned 117 Bloch-Wangsncss-Redfield theory 19911' Blocmbergen-Sorokin experiment - described 27511' - explained using spin temperature 28911' Boltzmilnn ratio Bil 255 Bond length, determined by magnetic resonance 82
Carr-Purcell sequence - explained 36711' - Meiboom-GilJ vllriely 36911' Chemical exchange, erred on NMlt 59211' Chemical shifts effect or electron spin 143f - experimental facts 88 - fonnal theory 92 Clebsch-Gordon coefficients 489 Coherence transfer, explained 331ff Collision broadening 213 Commutation relations for spins 18 Correlation fundion defined 193 - derived for a apecial case 584rr - general pN)pertietl 192ff Correlation time 193 COSY 3S0ff Cross-polarizlIlion (CP) - Harlmann-lIl1hll method 277lf - Pines-Gibby-Wllugh melhod 293fT Cross-relaxation and etllablishment of spin-lemperature 1<7 bet.... een ulllike nuclei 271ff Crystal fields - role in chemica.! shins 90,101 - role in electron spin resonance 50Sff Current density j(r), quantum mechanical expression for 93ff Curie's law, derived 163 Deeoupling of spins 303ff' Density malrix, su also Redfield theory of density matrix definition 159 general treatment lS7ff' interaction reprcsentlltion 163lf perturbation theory 164lf product operator method applied 344ff response to a 6-function 174lf in the rotating refcn:-nce frame 16Str for thermal equilibrium 162 time dependence using eigenfunctions 161
651
-
-
time-depend('nt equation 160 time development for l¥;'O spins 325ff of a two-I('...:I system 186ff used to derive transition probability
'OOff Detailed bahlllce 148 Diffusion effed of Carr-Pun:el1 pulse sequenc(' 367Fr eFred on resonance 367, 597ff Dirac equlltion, derivation of .-state coupling III Fr Double resonance 247ff classification of three kinds 247,248 - Cf"O!lSo-relaxation family 271tr - Pound·Overh ..user family 2Hff - spin-coherence family 2!l5ff S-flip only 296ff spin-echo variety (SEDOR) 311 Effective magnetic field, Ifelf , defined 21 I~leclric ficld gradient - asymmetry parameter, defined 495 calculated 5001f - introduced 486 - from valence eledrons 501 Eledric quadrul)()le (Ch..p.IO) 485ff - computation 500ff examples in strong and weak magnetic fields 497tr II..miltonian - for g('neral a.xes 496 - general treatment 486ff,49'11f - for principal aX('5 496 using raising al\d lowering operators 497 Sternhdmer IIntishielding factor 502 Elcctrk quadrul)()le mom(,llt, defined 496 Electron-nuclear double I"l'lIOnance (ENDOR) 2661f Electron spi n - coupling to nucleus l08lf - role in chemiclll shift 143tr - rol(' in Knight shift 113lf Electron apill echoes 524ff Electron spin I"l'lIOnl\nce (ESn) (Chap. II) 503ff Electrons electric field gradient produc<,d by 500 intNl\ction of electron orbit with nudei 89 interaction of electron slHn with nuclei 108ff magnetk interaction with nuclei (Chap.4) 87ff Equl'Ition of motion of a spin - classical, ill a rotating reference frame 12
652
-
Inductance, effect of X on 37 Interaction f'epre:lllentation 163f Irreducible tensor operators TI. M - applied to electron spin resonance 513f - construction from YI.M'S 493 defined -190 7iM'S 490 '11M 's ,193 use in electric quadrupole problems 4S9f Isomer shift 487
c1asskal theory II effect of all<'mating fields, c1l1l1Sical
20ff -
expeo:::tation value 17 quantum m«hllnical 13 in a rotating reference frame 29ff Ethyl alcohol 88 Exponential operators dcfined 25 expansion for cABe- A 459 a theorem I)f"oved 579,580 use in solving time-independent w;\\"e equation 26
J-ooopling 131f Jcener-Brockaert pulse sequence
Fcrmi function, defined 118 Ferromagnets, NMIt in 124lf Fourier transform (IT) NMR - derived 179ff - ('xperimental advantages 183,184 J-resolved 21) 323 line shapes in 20 32-1-5 two-dimellsionlll (2D), introduced 319ff I-rce induction d«ay, defined 23 {(...,), defined 71 Gau!e transfonnation, - defined 92 - invariance under 92f g-factor - defined 510 expression for 510 Lande 514 theory 505ff - for V k center 5:13ff g-shift su g-factor Gyromagnetic rl\tio "t, defined
defined 20 effect of finite 1/ 1 on time of the ecllO 49 effect of inhomogeneity on pl.llscs 422 relaled to power level 38,39 rell\tion to linearly polarized fl('ld 36 rotating 20 Ilartmann-irahn condition 277 I-jebel-Slichter equation for 1'1 150 Iligh-temperllture approximation - generl\1 discussion 589f - introduced and defined 63 1I)·periine structure (in ESR) - general thf!Ory 516f - in the v'" ~nter 533f
Magic angle - in the rotating frame (~Goldb"rg ('xperiment) 384ff - spinning 392ff Magic «hoes 388ff Magnetic dipoh.. r broa
380ff
Impedance, eff«t of X on Indirect nude"r coupling
38 131f
116
Larmor frequcncy 13 Laser 269f Line shapes - effect of crystalline anisotropy 605ff - thf!Ory of dipolar effed on 127ff I.inear respolllle thf!Ory applied to pulse NMR IHff Ill! basis for Fourier transform NMR 179ff - general thf!Ory 51 ff - introduced 51 Local field in double resonance 282 - in laboratory frame 222 - in robting frllme 235 Lorentz local field 125,182 Lorenlzian line, d('fined 39
11, -
380f
Kllight shift - experimental faclll 113 formuh. 122 - general theory 114f - relation to magnetization dellsity I
2
Magnetic dipolar terms A, B,C, D, E, P, defined 66 Magnetic field gradient defined 3SSf - effect wll('n diffU!lion is present 367, 369,597ff Magnetic field at nucleus - from elcctron orbit (c1assicl\l) 89 - from electron Sl)]n 108ff - quantum mC(;hanical 95 Magnetic resonance imaging (Mltl) 357ff Magu('tic susceptibility, X = X' - iX" - atomic theory of X' and X" 59lf complex x, introduced 37 - effect on coil impedance 37 further qmmtum mechanical expressiolls for SSOff mci\Sllremellt of sl)in contribution of conduction ele<:trons 122ff - orbital in terms of current density
produced by adiabatic demagnetization 227,2-45 Magnetic dipolar refocusing dil>olar coupling by «hoes 371ff
'0< - relation of X" to I(w) 71 fo,lagnetization, re[,..lion of rotatillg to linear polarizations 36 fo,lIIgnus ('xpansion 623f Maser 269f Master equation 149 Matrix el('menu of many-e\.ectron product functions 117 Meiboom-Gill method 369ff Mixer 184f Moments - effect of molccuh.r rotation 84 - further expressions for 582 - method of 71 - nth moment, (loin), delined 71 - lIC<:ond moment formulas 79,80 - ~eroth, firsl, and lICCond calculated
73ff
- (6...,")
delined 72 Motional narrowing - analyzed using the Bloch equations 592f - delillci1 213 - simple derivation 213f Multiple quantum coheN!:nce (01all.9) 4311f bll8ic cxplan"tioll 43-llf evolution, mixing, and detC(;tion 449-155 frequency selective pumping 434ff - generating a desired order 471ff - nOllsel«tive ~xcitation ...ol4lf - observin! a desired order ..63ff - preparation 445-449 - three or more spins ..55f
m(..) -
defined 51 derived by quantum lllC."Chani<:s
174ff
653
Nuclear Overhauser efTed (NOE) 257fT Nuclear polarization see Overhauser effect, Solid effed Operators set Exponential operators, Irreducible tensor operators, Raising and lowering operators Overhauser effect general theory 218ff _ nuclear (NOE) 257ff Perturbation theory, a useful theorem 585ff Phase coherent detection - apparatus 42 - and mixers 184f Powder line shapes, general thCQry 605ff Product operator method 341ff Pseudodipolar coupling 141 Pulse, 90 0 and 1800 , defined 22 Quadrature detection 185 Quadrupole see Electric quadrupole Quenching of orbital angular momentum 89ff Raising and lowering operators, introduced 14 Random phases, hypothesis applied to density matrix in thermal equilibrium 161 Rate phenomena, studied by NMR 592ff Redfield theory of the density matrix _ condition of validity 199 including applied alternating fields 215
including thermal equilibrium 204ff used to derive the Bloch equations 206f Redfield theory of saturation, see Spin temperature in the rotating frame Resonance, elementary quantum description 5ff Rotating reference frame - classical 12 - contrasted with interadion representation 168 _ doubly rotating frame 279ff _ quantum mechanical 29 Saturation - definition 8 - failure of classical saturation theory 231
Secular broadening 214 Sensitivity, the problem discussed
654
270
Signal·to-noise ratio and Carr-Purcell pulse sequences 367f and double resonance 270 _ and phase coherent detection 42 Single crystal spectra 127ff Slow motion, theory of effect on Tl~ 244ff Solid effect (in double resonance) 265 Spedral density 194 Spin echoes _ derivation using density matrix 169fT _ derivation using wave fundions 46fT - efTed of diffusion on 14, 597ff in ESR 524ff introduced (classically) 39fT in solids 371ff for spin I nuclei 60lff Spin-flip narrowing _ explained 406-409 general theory 409ff - observation 416ff - real pulses and sequences 421ff Spin-lattice relaxation Set also TI~ - in the Bloch equations 34 - general theory 145ff the Korringa law 15M in a metal 151fT when a spin temperature applies 146ff time, T1 , introduced 8 T 1 minimum 199 _ using the density matrix 197 Spin locking 245 Spinor 31fT Spin-orbit coupling - defined 504,505 - role in g-shift 505ff _ typical values 507 Spin susceptibility of condudion electrons defined 119 and Knight shift 119ff measured and theoretical values 1221f Spin temperature see also High-temperature approximation _ to analyze adiabatic and sudden changes 223ff _ in the laboratory frame 2211f _ in magnetism and magnetic resonance 219ff _ in the rotating frame (Redfield theory thereof) _ approach to equilibrium 239ff _ condition for validity 241 _ general theory 2341f _ inclusion of spin-lattice effects 242fT
~
~
TI
see Spin-lattice relaxation
used to analyze BloembcrgenSorokin experiments - used in 1'1 theory of solids 146ff Sternheimer antishielding factor 502
T"
- defined 243 - use for study of slow motions 2411f T2, introduced 31 Thermal equilibrium - density matrix corresponding to 163 - relation to transition rates 6lf Time-dependent Hamiltonians - corrections to appro:o;imate formulas 6231f - general theory 6161f Time development operator, defined 173
Time-ordering operator To (of Dyson), defined 617 Transition probability - formula and conditions of validity 61 - introduced Sf - role of the lattice in thermal processes
7.
Triangle rule 190 Two-dimensional Fourier transform method (20 IT) 319ft' .ret! also Fourier transform NMR Vk center
5331f
Wigner-Eekart theorem 489ff Wobble (in electron systems) 524 X(9), defined
17
655
Springer Series in Solid-State Sciences Editors: M. Cardona
P. Fulde
Prinriplelotl'llapdie H""""a.3rd Edition B)'C. P. Slichter 2 InlrodutliOflto Solid-State Theory 2nd Printing. By O. MaddulI& 3 D)"namial 5<'IUerin& or X-Ha)'S in Cl')'stals By Z. G. l'insker 4 Inelastic Elcdron Tunnelln& SpoectroKopy Editor: T. Wolfram 5 Fundamenlals or Cr)'stBI Grlmth I Macroscopic Equilibrium and Transport Concepts. 2nd I'rintin& By F. Rosenberger 6 Macnetit nux Slructures In Superconducton By R.P. tluebencr 7 <;""eo'l FUndiOftl in Qllanlunt 1'tIy:sia: 2nd Edition By E.N. Economou 8 SOlil_ ..... Condensed Maller 1'tI)'Sics 2nd Printint Edilon: A. R. Bi:Ulop and T. Schneider 9 P1NJlorerroeleetries By V. M. Fridkin 10 I'tIoeon DispeniolI Relatiolts-. Inwlaton By H. Bill: and W. Kre$$ II Electro. Tnnsporl in Compound Semieonducton By B. R. Nag 12 The I'h)"Sies or Elementary Enitalions By S. Nakajima, Y. Toyouwa. and R. Abe 13 'llIe l'h)·li.,. orSele"i"", and Tellurium Editors: E. Gerlach and I'. Grosse 14 Magnetit Bubble Technulogy 2nd Edition By A. H. Eschenfcldcr 15 Modem Cf)llallo&raphy I Symmetry of Crystals. Methods of Struaural Crystallography By B. K. Vainstuein 16 0'laoie Molenblr C.,stals Their Electronic: States. By E.A. Silinsh 17 n.eneo.yorMIfM1"m I St.alieland Dyl\ilmu. 2nd Edition By D.C. Mallis 18 ll.eluatiOfl of Elemutary EacitllinO'J Editon: R. Kubo and E. Hallamura 19 Solitons, Mathematical Methods for Physicists. 2nd Printin, By G. Eilenberger 20 Theory or Nonlinear Lanius 2nd Edition By M, Toda 21 Modem Crystallography II Structure of Crystals. By B. K. Vaill!iJnein. V, M, Fridkin. and V. L. Indenbom 22 )'111"1 Oerecls in Semlconduclors I Theoretical Aspects By M. lanooo and J. Bourgoin 23 Ph)'sies in On" Dimtllllot Editors: J. Bemasconi, T. Schneider 24 Pb,sia in Higb MallM:lie Fields Editors: S. Chlkuumi and N. Miura 2S FuodaJHnlll Pb,'Iia of AmOfJlhoDS S81Kondudon Editor. F. Yoneu.....
K. von Klitzing
H.·J. Queisser
26 Ebrstk l\IediIo willi Min05lr1lttare I One-DimcnWomJ Moods. By I.A. Kl.lnin n SuptffOlldudivily or Tnllliilioll Melals Their Alloys and Compollnds.
By S. V. VOll5OYSky. Yu.A. Izyumol', and E.Z. Kurmaev 28 The SlrliClurt and Pro~rlil's of Matter Editor: T. Matsubara 29 ElettrOIl Correlatlollind Magnetism in Narrow-nand 5)'SleD's Editor: T. MOTi)'. 3(1
Slalisl;ul Physics I 2nd Edition
By M. TolIa, R. Kubo. N. Saito 3\ Statistical Ph)'skt II By R. Kubo. M. TOOa. N. Hashilsume 32 QlllInlllID Theory of ~bl"e'im1 By R.M. White )J Mix"' CI')'UIlb: By A.I. Kilaipodsky ~ i>ta.--: ~Of)' and upecrimUlli I Lanitt DyIWltics and Models of InlCrlllomM; Fom:s. By P. Bnicsdl 35 PoinT ~r_ In SftnltolKlDdon II E:
36 Mock'm Cr)~.1IoI ...pby III Crystal Gm'O,lh ~nd Edition ByA.A.C1Icrnov 37
~1odem
C..,-staJJoXnphy IV
Physical Properties of Crystals By L.A. Shuvalov
3S Ph)'sics of Inlertalarlon Compounds Editors: L. Pictronero and E. Tosall; 39 Andenon ~Jiution Editors: Y. NagllOkaand H. Fukuyama ~ SemKonductor Ph)'sKs An Introduction 3rd Edition By K. Secaer 'II n.e urro Method Muffin-Tin Orbilals and Btttronic Struaure ByH.LSk~r
42 Cf)'SlII Opfla ..flh Spatlal Dlspenioa, lid Ex
l